{ "0011/astro-ph0011134_arXiv.txt": { "abstract": "Ground-based \\& HST images of the nearby galaxy SBS~1543+593 ($z=0.009$) show it to be a Low Surface Brightness (LSB) galaxy with a central surface brightness of $\\mu_B(0)\\:=23.2$ mag arcsec$^{-2}$ and scale length 0.9~\\h , values typical for the local LSB galaxy population. The galaxy lies directly in front of the QSO HS~1543+5921 ($z=0.807$); an HST STIS spectrum of the quasar reveals a damped \\lya\\ (D\\lya ) line at the redshift of the interloper with an H~I column density of $\\log N$(H~I)$\\:=\\:20.35$, as well as several low-ionization metal lines with strengths similar to those found in the Milky Way interstellar medium. Our data show that LSB galaxies are certainly able to produce the D\\lya\\ lines seen at higher redshift, and fuels the speculation that LSB galaxies are a major contributor to that population of absorbers. ", "introduction": "In the course of spectroscopic follow-up observations of quasar candidates in the Hamburg Quasar Survey, a redshift of $z=0.807$ was measured for the QSO HS~1543+5921. However, Schmidt-plate images showed the object to be very extended, suggesting that the QSO might be centered on a low redshift galaxy. Follow-up observations by \\citet[hereafter RH98]{RH98} found the QSO to be aligned with the foreground galaxy SBS~1543+593, and measurement of an H~II region in a spiral arm revealed the galaxy to be at $z=0.009$. Although detecting QSOs close to nearby galaxies is not difficult, finding one which shines through the center of a galaxy and which is bright enough to be observed spectroscopically with HST is extremely rare. In this paper we present ground-based and HST optical images of SBS~1543+593 and show that the galaxy is actually a Low Surface Brightness (LSB) galaxy. We also present an HST spectrum of the background QSO HS~1543+5921, which reveals a damped \\lya\\ (D\\lya) absorption line at the redshift of the foreground galaxy. ", "conclusions": "Our optical and spectroscopic data have shown that SBS~1543+593 is a LSB galaxy causing a D\\lya\\ system at $z=0.009$ in the spectrum of HS~1543+5921. This makes it the lowest redshift D\\lya\\ system discovered outside of the local group. The identification of D\\lya\\ from a {\\it known} LSB is important, for the following reasons. So far, results from ground-based and HST imaging of fields around QSOs known to show $z<1$ D\\lya\\ lines have been surprising, with the detection of a whole variety of galaxy types, including normal early and late-type HSB spirals, and amorphous LSB galaxies, identified as responsible for the absorption \\citep{steidel95, steidel94, Lanz97, LeBr97, RaTu98, pett00}. This wide variety of absorber types has led to the speculation that D\\lya\\ systems may not simply signal the presence of normal gas-rich spiral galaxies after all, as has been postulated since their discovery \\citep{wolfe86}. It is important to note, however, that in most cases there exists no redshift information for the purported absorbers. Proximity to the line of sight is no guarantee that an `identified' object is the absorber, since there are often several absorption systems at redshifts other than that of the D\\lya\\ line along the QSO line of sight. If galaxies are responsible for these other systems, then the chance of mis-identification is high (particularly if D\\lya\\ systems do not arise in normal galaxies). Hence the detection of a D\\lya\\ line from SBS~1543+593 is unique, in that we know unequivocally that the absorber is an LSB galaxy. It also seems likely that if SBS~1543+593 were moved to a redshift similar to those of the $z<1$ D\\lya\\ systems already studied, it would be extremely difficult to detect, partly due to its low surface brightness and partly due to its close proximity to the QSO and its small angular size. In fact, in this case, it is more likely that the nearby barred spiral galaxy south-west of the pair, which we take to be background to SBS~1543+593, would be identified as the D\\lya\\ absorber if no redshift information were available. Our observations of SBS~1543+593 support the idea that LSB galaxies may contribute significantly to the population of D\\lya\\ absorbers, as initially suggested by \\citet{Imp89}. Although finding a D\\lya\\ from an LSB galaxy does not prove that {\\it all} D\\lya\\ systems are LSB galaxies, our detection does prove that LSB galaxies can produce such systems. Perhaps more significant is the detection of this QSO-galaxy pair in the first place. Finding any bright QSO shining through the center of a nearby galaxy is extremely rare, and it is intriguing that an LSB galaxy is the interloper. LSB galaxies are believed to be relatively free of dust, and hence optically thin to the photons of background objects --- these may therefore be the best types of galaxies to allow the light of quasars to shine through. This potential selection effect has been a concern for interpreting the copious abundance measurements of high-redshift D\\lya\\ systems: if a particular type of galaxy preferentially favors QSO light passing through it, then the derived metallicities are only applicable for that type of galaxy. There are also theoretical bases for the idea that LSB galaxies hold most of the high column density H~I cross-section: for example, \\citet{raul99} predicted that HSB disks consume neutral gas too fast to explain the observed evolution in the neutral gas mass density with redshift, and that consumption of hydrogen by LSB galaxies better fits the abundance measurements. The mere existence of the QSO-galaxy pair discussed herein seems to strongly support these views. Unfortunately, just how common such QSO/LSB-galaxy alignments are at higher redshift will always be difficult to determine, due to the intrinsic faintness of the galaxies, especially if the galaxy is close to the QSO sightline as with this pair. Finally, we note that HS~1543+5921 is bright enough to be used to measure abundances in the interstellar medium of SBS~1543+593 with further HST observations. The current data do not have sufficient resolution for reliable metallicity estimates, so follow-up UV observations would be extremely valuable. Comparison of gas metallicities from a known LSB galaxy with higher redshift D\\lya\\ systems would be useful in deciding if LSB galaxies are, in general, responsible for the absorption. Measurement of the alpha-to-iron elemental abundances in SBS~1543+593 could also be used to derive the star-formation history of the galaxy, and relative metal abundances may provide insight into its dust content. Lastly, comparison of the kinematic structure of metal lines arising from SBS~1543+593 could be matched to global H~I kinematics derived from 21~cm emission maps of the galaxy, and to profiles seen in high-redshift D\\lya\\ lines which are believed by some authors to be indicative of rotating disks." }, "0011/astro-ph0011072_arXiv.txt": { "abstract": "We present here numerical hydrodynamic simulations of line-driven accretion disk winds in cataclysmic variable systems. We calculate wind mass-loss rate, terminal velocities, and line profiles for \\ion{C}{4} (1550~\\AA) for various viewing angles. The models are 2.5-dimensional, include an energy balance condition, and calculate the radiation field as a function of position near an optically thick accretion disk. The model results show that centrifugal forces produce collisions of streamlines in the disk wind which in turn generate an enhanced density region, underlining the necessity of two dimensional calculations where these forces may be represented. For disk luminosity $L_{disk}=L_{\\sun}$, white dwarf mass $M_{wd}=0.6M_{\\sun}$, and white dwarf radii $R_{wd}=0.01R_{\\sun}$, we obtain a wind mass-loss rate of $\\dot M_{wind}=8 \\times 10^{-12} M_{\\sun} {\\rm yr}^{-1}$, and a terminal velocity of $\\sim 3000 {\\rm \\; km \\; s}^{-1}$. The line profiles we obtain are consistent with observations in their general form, in particular in the maximum absorption at roughly half the terminal velocity for the blue-shifted component, in the magnitudes of the wind velocities implied by the absorption components, in the FWHM of the emission components, and in the strong dependence in inclination angle. ", "introduction": "The first evidence for winds from cataclysmic variables (CVs) came from the discovery of P~Cygni profiles in the UV resonance lines of SS~Cygni (\\cite{hea85}). The dependence on the observability of CV winds with inclination angle and the apparent similarity between cataclysmic variable line profiles and those of OB stars led to the early suggestion by C\\'ordova \\& Mason (1982) that the winds in CVs originate from the accretion disk and that the line radiation pressure is responsible for the wind. More recently, P~Cygni profiles have been detected from virtually all nonmagnetic CVs with high mass accretion rates ($\\gtrsim 4\\times 10^{-10} M_{\\sun} {\\rm yr}^{-1}$) (e.g., \\cite{cor82}; \\cite{gre82}; \\cite{gui82}; \\cite{pri95}; \\cite{fri97}; \\cite{gan97}; \\cite{kni97}). The absorption component is most apparent in low inclination systems, and is not detected in eclipsing systems (e.g., \\cite{hut80}; \\cite{kra81}; \\cite{hol82}; \\cite{cor85}; \\cite{mas95}; \\cite{che97}). Past efforts in the development of models for CV winds focused on one-dimensional models (\\cite{vit88}; \\cite{kal88}), and on kinematic modeling (\\cite{shl93}; \\cite{kni95}), which succeeded in showing consistency between the assumed polar geometry of a disk wind and observed profiles. Icke~(1980) developed a two-dimensional disk wind model but did not take into account the radiation pressure due to line absorption. For a typical white dwarf mass of $0.6M_{\\sun}$ (\\cite{lei80}) the disk luminosity required to produce such a wind without line radiation pressure (assuming radiation pressure due to continuum scattering only) would require a luminosity of $\\sim 10^5 L_{\\sun}$ (\\cite{ick80}), several magnitudes above the observed luminosity of CVs which vary between $0.01L_{\\sun}$ and $10 L_{\\sun}$ (\\cite{pat84}). Icke did find that for accretion disk winds in general, for sufficiently high disk luminosity, a biconical disk wind would be produced. Icke~(1981) suggested that biconical winds were a general property of accretion disk winds independent of the wind driving mechanism. Murray et~al. (1995) found that for line-driven disk winds in active galactic nuclei (AGN) the wind flow tends to be parallel to the accretion disk and developed a one-dimensional disk wind model for these systems. Murray \\& Chiang (1996) found that the single-peaked optical emission lines seen in AGNs and high-luminosity CVs could be accounted for by emission at the base of an accretion disk wind. Using the disk wind model of Murray et~al. (1995), Murray \\& Chiang (1997) calculated synthetic line profiles of \\ion{C}{4} 1550\\AA \\ and found qualitative agreement between their models and observed line profiles. In the case of CVs, as we discussed above, the winds are most apparent in low-inclination systems and not observed in eclipsing systems, indicating that the winds in CVs tend to flow perpendicular to the accretion disk. In an earlier paper (\\cite{per97}, hereafter Paper~I) we presented two-dimensional isothermal hydrodynamic models of line-driven accretion disk winds (LDADW) in CVs. To our knowledge these isothermal models were the first two-dimensional hydrodynamic models developed for LDADWs. Results from Paper~I show, in analogy with line-driven winds from early type stars, that terminal velocities are approximately independent of the luminosity of the disk, although increments in luminosity produce increments in mass-loss rate. In Paper~I we showed that rotational forces are important in the study of winds from accretion disks, and that they cause the velocity streamlines to collide. The collision of streamlines reduce the speed and increase the density of the wind producing an enhanced density region. In Paper~I we also showed that the highest absorption occurs in the enhanced density region where density is increased relative to a spherically diverging wind with the same mass loss rate and the velocity is roughly half the terminal velocity. This density increase is necessary in order to produce at least marginally optically thick lines. Recently Proga, Stone, \\& Drew (1998) also developed two-dimensional models for these systems, obtaining similar results to the models presented here. When evaluating the line-radiation pressure from the disk, in order to make the computations manageable, they assumed that the velocity gradient is primarily along the ``z'' direction, as we also did in Paper~1 and in the models presented here. Proga, Stone, \\& Drew worked out their hydrodynamic model in spherical coordinates, while the models presented here were implemented in cylindrical coordinates. They found similar wind velocities and wind mass loss rates as we do here, although they assumed disk luminosities about an order of magnitude greater. A difference we find with the results of the models presented by Proga, Stone \\& Drew (1998) is that they obtain unsteady flows characterized by large amplitude fluctuations in velocity and density, while for the models presented in this paper we find a steady flow. A shortcoming of our previous work, presented in Paper~I, was that we had assumed an isothermal wind and an isothermal disk. In this work we develop a hydrodynamic model of LDADWs in CVs which includes the radial structure of an optically thick accretion disk with the corresponding radiation fields and surface temperature distributions. We have also implemented the corresponding energy conservation equation for the models presented here self consistently (see equation~[18]), including the adiabatic heating and cooling effects due to compression and expansion. For the models presented in this paper we have also taken values presented by Abbott~(1982) for the line radiation force multiplier parameters, rather than the earlier values used by Castor, Abbott, \\& Klein~(1975) as we did in our isothermal models presented in Paper~I. From the results of the hydrodynamic model presented in this paper, we calculate theoretical line profiles for the \\ion{C}{4} 1550~\\AA \\ line, and find that the line profiles obtained through our model are consistent with observations in their general form and strong dependence with inclination angle. Thus we are able to predict the general observed wind properties of CVs. The boundary layer where the disk intercepts the white dwarf and the white dwarf photosphere may also contribute to the radiation field. The white dwarf by itself (without considering the effects of the accreting mass) will have a luminosity of the order of $0.01L_{\\sun}$(\\cite{lei80}), while the accretion disk will have typical luminosities of the order of $L_{\\sun}$. In a steady state accretion disk half of the energy of the accreting mass is converting into radiation emitted by the accretion disk. Assuming a relatively slowly rotating compact star, it has been predicted that the boundary layer presents a luminosity approximately equal to the accretion disk, thus accounting for the other half of the energy of the accreting mass (e.g., \\cite{fra92}). But there is still uncertainty concerning the existence and spectrum of the boundary layer (\\cite{vrt94}; \\cite{mau95}; \\cite{lon96}). Analysis of observations have found, in some cases, a boundary layer luminosity an order of magnitude less than the disk luminosity (e.g., \\cite{mau95}), and in other cases a boundary layer luminosity comparable to disk luminosity (\\cite{lon96}). In the models presented here we have assumed that the radiation field is generated by the accretion disk alone and have therefore neglected any boundary layer or white dwarf radiation. In the future we plan to explore the effects of boundary layer radiation on our models. In \\S~2 we discuss the radial structure of the accretion disk and the radiation field as implemented in our model. In \\S~3 we derive the expression we use for the treatment of the line radiation pressure. In \\S~4 we present and discuss the hydrodynamic calculations. In \\S~5 we present and discuss theoretical \\ion{C}{4} 1550~\\AA \\ line profiles and compare them with observations. In \\S~6 we present a summary and conclusions of this work. ", "conclusions": "We have developed a 2.5-dimensional hydrodynamic line-driven adiabatic accretion disk wind model. Our model solves a complete set of adiabatic hydrodynamic partial differential equations, using the PPM numerical scheme and implementing the radial temperature and radiation emission distributions on the surface of an accretion disk. From our models we calculate wind mass-loss rates, terminal velocities, and line profiles for \\ion{C}{4} (1550~\\AA) for various angles. For typical cataclysmic variable parameter values of disk luminosity $L_{disk}=L_{\\sun}$, white dwarf mass $M_{wd}=0.6M_{\\sun}$, and white dwarf radii $R_{wd}=0.01R_{\\sun}$, we obtain a wind mass-loss rate of $\\dot M_{wind}=8 \\times 10^{-12} M_{\\sun} {\\rm yr}^{-1}$, and a terminal velocity of $\\sim 3000 {\\rm \\;km \\;s}^{-1}$. Studies of observed ultraviolet line profiles from CVs have estimated CV wind mass loss rates to be in the order of $10^{-12}$ to $10^{-11} M_{\\sun} {\\rm yr}^{-1}$ and found CV wind terminal speeds in the order of $\\sim 3000 \\; {\\rm km} \\; {\\rm s}^{-1}$ (\\cite{kra81}; \\cite{cor82}; \\cite{gre82}; \\cite{mau87}; \\cite{pri95}; \\cite{fri97}), thus the models developed in this work predict values for wind mass loss rates and terminal speed consistent with observations. Furthermore, the line profiles we obtain through our computational model are also consistent with observations in their general form, in particular in the maximum absorption at roughly half the terminal velocity for the blue-shifted component, in the magnitudes of the wind velocities implied by the absorption components, in the FWHM of the emission components, and in the strong dependence in inclination angle. We have shown that rotational forces are important in the study of winds from accretion disks. They cause the velocity streamlines to collide which results in an enhanced density region. In this region the wind speed is reduced and the wind density is increased. The increase in density caused by the collision of streamlines is important because it permits the appearance of blue-shifted absorption lines as observed in P~Cygni profiles of low-inclination CVs. A shortcoming of our models is the neglect of radiative cooling and heating and local ionization equilibrium throughout the wind. In the next paper of this series we will present models which include these effects. We also plan to extend our models to the study of the wind dynamics of low mass X-ray binaries and active galactic nuclei, where these processes may play a more important role in determining the overall dynamics." }, "0011/astro-ph0011244_arXiv.txt": { "abstract": "{ We present the results of a study based on an \\XMM\\ Performance Verification observation of the central 30\\arcmin\\ of the nearby spiral galaxy M31. In the 34-ks European Photon Imaging Camera (EPIC) exposure, we detect 116 sources down to a limiting luminosity of \\scinot{6}{35}~\\Lcgs\\ (0.3--12~keV, $d=760$~kpc). The luminosity distribution of the sources detected with \\XMM\\ flattens at luminosities below $\\sim2.5 \\times 10^{37}$~\\Lcgs. We make use of hardness ratios for the detected sources in order to distinguish between classes of objects such as super-soft sources and intrinsically hard or highly absorbed sources. We demonstrate that the spectrum of the unresolved emission in the bulge of M31 contains a soft excess which can be fitted with a $\\sim$0.35-keV optically-thin thermal-plasma component clearly distinct from the composite point-source spectrum. We suggest that this may represent diffuse gas in the centre of M31, and we illustrate its extent in a wavelet-deconvolved image. ", "introduction": "\\begin{figure*} \\vbox{\\psfig{figure=fig1half.ps,width=12cm}\\vspace{-8.0cm}} \\hfill\\parbox[b]{5.5cm}{ \\caption{Three-colour \\XMM\\ EPIC MOS1 image of the central 30\\arcmin$\\times$30\\arcmin\\ of M31. The red, green, and blue intensities correspond to logarithmically-scaled counts from energy bands of 0.3--0.8~keV, 0.8--2~keV, and 2--8~keV respectively. The image was constructed with 2\\arcsec\\ pixels and has been smoothed with a Gaussian of FWHM 4.4\\arcsec, approximately equal to that of the MOS1 PSF. \\vspace{4.0cm} } } \\label{fig:image} \\end{figure*} Being the closest spiral galaxy to our own, the Andromeda Galaxy (M31) is in many respects ideal for the study of X-ray emission in a galaxy similar to the Milky Way. The sources in M31 are observed at a nearly uniform distance, and, owing to the inclination of the galaxy (77$^{\\circ}$), they are viewed through a substantially lower absorption column than for sources in the Galactic plane. In a recent review, van den Bergh (2000) reports a distance modulus to M31 of 24.4$\\pm$0.1, corresponding to a distance of 760~kpc. We adopt this value in our analysis, and for consistency we scale to this distance when discussing published luminosities which assume a different distance. Over 100 X-ray sources in M31 were detected with the \\Einstein\\ observatory (Trinchieri \\& Fabbiano 1991; van Speybroeck et al.\\ 1979). The brightest X-ray source in M31 was found to have a luminosity of $\\sim\\mscinot{3}{38}$~\\Lcgs, approximately the Eddington luminosity for spherical accretion onto a 1.4~\\msun\\ neutron star. For sources down to $\\mscinot{2}{36}$~\\Lcgs, the luminosity distribution was reported to be consistent with a single power law which, extrapolated to fainter levels, could fully account for the X-ray emission from the bulge of M31. Primini, Forman \\& Jones (1993) detected 86 X-ray sources in the central 34\\arcmin\\ of M31 with the \\ROSAT\\ HRI. They found a break in the luminosity distribution at $\\sim\\mscinot{2}{37}$~\\Lcgs, below which the distribution of sources flattened. This flattening suggested that the detected population of X-ray sources could account for only $\\sim$15--26\\% of the unresolved X-ray emission in M31. Contributions from known less-luminous populations of X-ray sources also could not fully account for the unresolved emission, suggesting that the remaining emission is truly diffuse or due to a new class of X-ray sources. In an extensive, 6.3 deg$^2$, survey of M31 with the \\ROSAT\\ PSPC, 396 X-ray sources were detected (Supper et al.\\ 1997). However, only 22 of these sources were detected in the bulge region ($r < 5$\\arcmin) due to the resolution of the PSPC. In the first \\Chandra\\ observation of M31, the nuclear source seen with \\Einstein\\ and \\ROSAT\\ was resolved into five sources (Garcia et al.\\ 2000). One of these sources is located within 1\\arcsec\\ of the radio nucleus of M31 and exhibits an unusually soft X-ray spectrum, suggesting that it may be associated with the central super-massive black hole. A few more pairs of previously unresolved sources and a new transient were also detected within 30\\arcsec\\ of the nucleus. M31 was selected as an \\XMM\\ (Jansen et al.\\ 2001) Performance Verification target in order to demonstrate the capabilities of \\xmm\\ in performing spectral and timing studies in a field of point sources and extended emission. In this paper we focus on the group properties of the X-ray point sources in M31 as well as the diffuse emission. In a companion paper (Osborne et al.\\ 20001, Paper II), we discuss the spectral and timing properties of individual X-ray sources in M31. ", "conclusions": "Using \\XMM\\ data, we have confirmed that the X-ray emission from the bulge of M31 is dominated by bright point-like sources, most of which are likely to be low-mass X-ray binaries. For sources in the central region of M31, we have confirmed that the luminosity distribution is flatter toward lower luminosities (Primini et al.\\ 1993). The steepening of the luminosity distribution above $2.5 \\times 10^{37}$\\Lcgs\\ is indicative of a lack of bright sources in M31 (cf.\\ the source distribution in M33, Long et al.\\ 1996). Only two sources in our sample have a $0.3$--$12$ keV unabsorbed luminosity $\\geq 10^{38}$\\Lcgs. As in previous observations with \\Einstein\\ and \\ROSAT\\ (Trinchieri \\& Fabbiano 1991; Primini et al.\\ 1993), significant unresolved emission was found to contribute to the total emission of the bulge. The flattening of the luminosity distribution for fainter sources means that an extrapolation of the detected population of point sources at lower energies cannot account for the total core emission of M31. A soft excess in the spectrum of the M31 bulge was previously reported in \\ROSAT\\ and \\BeppoSAX\\ observations (Irwin \\& Bregman 1999; Trinchieri et al.\\ 1999). Our analysis of the \\XMM\\ data shows that the soft component in the spectrum of the bulge is associated with unresolved emission; this confirms the results of Borozdin \\& Priedhorsky (2000) based on \\ROSAT\\ data. More importantly, our \\XMM\\ study has revealed for the first time that while the integrated spectrum of point-like sources is featureless, the spectrum of the unresolved emission shows multiple emission lines typically found in the spectrum of hot, optically thin plasma. Therefore, we suggest that the second significant source of X-ray emission in the bulge is truly diffuse gas with an effective temperature $\\sim$0.35 keV. The contribution of this gas to the total unabsorbed X-ray luminosity is estimated to be $\\sim$10\\% in 0.3--12~keV band (corresponding to $\\approx 2 \\times 10^{38}$\\Lcgs), but more than 20\\% in the \\ROSAT\\ band (0.1--2.4~keV). The significance of this result goes far beyond the case of M31, because the bulge of this galaxy is often considered as a prototype for the population of early-type X-ray galaxies. For example, Sarazin et al.~(2000) recently reported that, according to \\Chandra\\ observations, 23\\% of X-ray emission from NGC\\,4697 is emitted by interstellar gas, contrary to their previous expectations (Irwin et al.\\ 2000). Two more \\XMM\\ observations of the central region of M31 are scheduled as part of the Guaranteed-Time program, as are observations of five additional fields along the disk of M31. These will allow us to reach fainter flux levels in the bulge and to study the populations of X-ray sources in different parts of the M31 galaxy." }, "0011/hep-ph0011353.txt": { "abstract": "We review new results in neutrino physics, including the latest data of the Super-Kamiokande, SNO and K2K experiments. ", "introduction": "The recent evidence for neutrino oscillations \\cite{SK1} provided us with the first firm evidence of physics beyond the standard model and opened a new an exciting era in neutrino studies. Neutrino physics is a very active branch of particle physics now, both experimentally and theoretically. The experimental data keep pouring in, and new experiments are either under way or in an advanced stage of planning. On the theoretical side, there are new analyses of the data, both in the 3-flavour and 4-flavour schemes; the analyses of the solar neutrino data are being extended to cover the ``dark side'' of the parameter space, not studied (or little studied) before; the reach of future planned experiments is being investigated and new experiments designed to eliminate the white spots on the map of neutrino properties are being suggested. In addition, since the discovery of neutrino oscillations a large number of models of neutrino mass was proposed and some old models were reconsidered in the light of the new data. In the present lectures we review new results in neutrino physics, mainly the experimental ones. We discuss the latest results of the solar and atmospheric neutrino experiments and their analyses; results from the reactor and accelerator experiments, including those from the first long-baseline accelerator neutrino experiment K2K, and then briefly discuss future experiments and projects. Finally, we discuss how all the presently available data can be summarized concisely in terms of the phenomenologically allowed structures of the neutrino mass matrix. We do not discuss neutrino mass model because of the lack of space; for recent reviews on this subject the reader is referred to \\cite{Sm1,AltFer,Tan,Moh1,Fr,Barr1}. The data reviewed in the lectures given at the NATO ASI in Cascais included the results reported at the Neutrino 2000 Conference in Sudbury, June 16 - 21, 2000. However, the present written version is updated to include the results reported up to November of 2000. %up to and including those reported at the ICHEP in Osaka, July 27 - August %2, 2000. For other recent reviews on neutrino physics see, e.g., \\cite{Bil,Akh1}. ", "conclusions": "" }, "0011/astro-ph0011522_arXiv.txt": { "abstract": "We combine our compilation of photometry of M31 globular clusters and probable cluster candidates with new near-infrared photometry for 30 objects. Using these data we determine the globular cluster luminosity function (GCLF) in multiple filters for the M31 halo clusters. We find a GCLF peak and dispersion $V_0^0=16.84\\pm0.11$, ${\\sigma}_t=0.93\\pm0.13$ (Gaussian ${\\sigma}=1.20\\pm0.14$), consistent with previous results. The halo GCLF peak colors (e.g. $B_0^0-V_0^0$) are consistent with the average cluster colors. We also measure $V$-band GCLF parameters for several other subsamples of the M31 globular cluster population. The inner third of the clusters have a GCLF peak significantly brighter than that of the outer clusters ($\\Delta V^0\\approx 0.5$~mag). Dividing the sample by both galactocentric distance and metallicity, we find that the GCLF also varies with metallicity, as the metal-poor clusters are on average 0.36~mag fainter than the metal-rich clusters. Our modeling of the catalog selection effects suggests that they are not the cause of the measured GCLF differences, but a more-complete, less-contaminated M31 cluster catalog is required for confirmation. Our results imply that dynamical destruction is not the only factor causing variation in the M31 GCLF: metallicity, age, and cluster initial mass function may also be important. ", "introduction": "Globular clusters (GCs) are unique markers of the formation and evolution of galaxies. They are bright, easily identifiable packages of Population~II stars with homogeneous abundances and history. GCs are relics of the earliest star formation in galaxies and have witnessed many changes in their parent galaxies. The study of globular cluster systems (GCSs) is thus important in understanding the history of galaxies. The distribution of integrated GC magnitudes, known as the globular cluster luminosity function (GCLF), has long been known to be unimodal and approximately symmetric in the Milky Way. The assumption that these properties are universal has allowed the determination of GCLF parameters for over a hundred other galaxies, and the peak absolute magnitude is found to be roughly constant from galaxy to galaxy \\citep{har88}. Since the variation in mass-to-light ratios among GCs is fairly small \\citep[see, e.g.,][]{dg97} the constant peak magnitude implies the existence of a characteristic mass scale. Theorists have attempted to explain why there should be a characteristic mass scale for globular clusters: is it a property of formation \\citep[e.g.][]{pd68}, a result of subsequent dynamical processes \\citep[e.g.][]{ot95}, or a combination? The constant peak magnitude also presents a challenge to observers, who have attempted to quantify the variation in peak magnitude by environment \\citep{bt96}, galaxy type and luminosity \\citep{whi97}, and color \\citep{kws99}. Many observers have also attempted to use the GCLF peak as a standard candle for distance measurement. This method has had a mixed reception, with some authors \\citep[e.g.][]{whi97} claiming good results and others \\citep[e.g.][]{fer99} less complimentary. GCLF measurements are made in a number of different bandpasses. The most widely used are $V$ and $B$, (and their Hubble Space Telescope (HST) equivalents, F555W and F450W); the I-band equivalent F814W is also commonly used for HST data. $B$ and $V$ band mass-to-light ratios are sensitive to metallicity, so GCSs with the same underlying mass distribution but different average metallicities will have different GCLFs. \\citet*{acz95} used the relation between metallicity and mass-to-light ratio as predicted by \\citeauthor{w94}'s (1994) simple stellar population models to estimate the effects of metallicity variations on the GCLF peak. Metal-rich clusters should be fainter (their $M/L$ is larger), and the effect on the GCLF can be substantial. For example, \\citet{acz95} predict that a change in mean GCS metallicity from $\\overline{\\rm [Fe/H]}=-1.35$~dex (the Milky Way value) to $-0.60$~dex \\citep[the value for the elliptical NGC~3923;][]{z95} shifts the GCLF peak by $\\Delta B^0=0.35$ and $\\Delta V^0=0.22$. The same metallicity change shifts the $J$ band GCLF peak by $\\Delta J^0=-0.09$, because the metallicity effect on $M/L$ changes direction for bandpasses redward of $I$. Observing at longer wavelengths also reduces the effects of extinction, both Galactic and within the GCS parent galaxy. Destruction of globular clusters through dynamical effects could produce a globular cluster mass function (GCMF) that varies with distance from the center of the host galaxy. The three effects usually considered by modelers are disk shocking, evaporation, and dynamical friction; the first two destroy low-mass, low-concentration clusters, while the last is most effective for massive clusters. All three effects are strongest at small Galactocentric distance $R_{gc}$. Even at small $R_{gc}$, the dynamical friction timescale for typical-mass GCs is much longer than a Hubble time, and dynamical friction is probably not important in GCMF evolution \\citep{ot95}. Low-mass clusters near the Galactic center are thus most prone to destruction. Assuming a mass-to-light ratio independent of $R_{gc}$ -- if the GCS has a radial metallicity gradient this is not strictly true since $M/L$ depends on [Fe/H] -- the radial difference in the GCMF can be translated into a radial difference in the GCLF. Many authors \\citep{b97,og97,v98} have predicted the size of the GCLF variation, with widely differing results (see Section~\\ref{sec-implic}). In \\citet{b00} we compiled the best available photometry for 435 M31 globular clusters and plausible candidates. This catalog contains $V$ magnitudes for almost all objects and $B$ magnitudes for about 90\\%, but completeness in the longer-wavelength bands is much lower. About 55\\% of the objects have $I$ magnitudes, and the same fraction, although not the same objects, have $JK$; Figure~\\ref{magcomp-ir} shows the completeness of the existing IR photometry. In this paper we estimate the parameters of the halo clusters' GCLF in six different bandpasses; this includes the first GCLF measurements in the near-IR. We then investigate the variation in the $V$-band GCLF parameters for several different subsamples of M31 GCs. Finally, we consider the implications of the measured GCLF variations for GCS and galaxy formation and evolution. ", "conclusions": "We have calculated the first $URJK$ GCLFs for M31 halo globular clusters, and find that the GCLF peak colors are consistent with the average cluster colors. Our parameters for the $V$- and $B$-band halo GCLFs are consistent with those of other groups. We find no significant differences between the disk and halo GCLF peaks, although the disk has a lower dispersion. A difference in GCLF peak at the $2\\sigma$ level occurs when we consider the inner and outer-most groups, as determined by projected galactocentric distance. This difference is consistent with that predicted by \\citet{og97} for M31; however, we do not detect the predicted difference in GCLF dispersion. We separate the M31 clusters by metallicity and find that the metal-rich clusters have a brighter GCLF peak than the metal-poor clusters, even when the difference in $R_{gc}$ is taken into account. Modeling of the catalog selection effects suggests that these effects are not responsible for the measured GCLF differences. However, an M31 GC catalog with well-understood and spatially uniform completeness and contamination is required in order to definitively confirm our results. Such a catalog might be produced by a near-IR, high spatial resolution survey of M31. We consider the implications of the GCLF differences for models of globular cluster and GCS formation, and conclude that younger ages for metal-rich clusters plus dynamical destruction of inner clusters are the most likely causes of the observed GCLF variations." }, "0011/astro-ph0011008_arXiv.txt": { "abstract": "We present optical spectroscopic identification of sources identified in the BeppoSAX MECS fields of the High Energy LLarge Area Survey (HELLAS). In total 62 sources from a sample of 115 brighter than $F_{5-10keV}>5\\times10^{-14}$ $cgs$ have been identified. We find a density of 13-21 sources/deg$^2$ at $F_{5-10keV}>5\\times10^{-14}$ $cgs$, which contribute to $20-30$ \\% of the hard X-ray background. Evidences are found for type 1 AGN being more absorbed with increasing redshift or luminosity. The low redshift ($z<0.2$) ratio of type 2 to type 1 AGN is $3\\pm1.5$, in agreement with the unified models for AGN. The luminosity function of type 1 AGN in the 2-10 keV band is preliminary fitted by a two-power law function evolving according to a pure luminosity evolution model: $L\\propto (1+z)^{2.2}$. ", "introduction": "Hard X-ray observations are the most efficient way of tracing emission due to accretion mechanisms, such in Active Galactic Nuclei (AGN), and sensitive hard X-ray surveys are powerful tools to select large samples of AGN less biased against absorption and extinction. In this framework we decided to take advantage of the large field of view and good sensitivity of the BeppoSAX MECS instrument (Boella et al. 1997a,b) to survey tens to hundreds of square degrees at fluxes $\\ga 5-10\\times10^{-14}~ erg~cm^{-2}s^{-1}$ (Fiore et al. 2000a), and using higher sensitivity XMM-Newton and Chandra observations to extend the survey down to $\\sim10^{-14}~ erg~cm^{-2}s^{-1}$ on several deg$^2$. The results from the optical identification of a sample of faint Chandra sources discovered over the first 0.14 deg$^2$ have been published by Fiore et al. (2000b). This approach is complementary to deep pencil beam surveys ($\\sim0.1$ deg$^2$, see e.g. Mushotzky et al. 2000, Hornschemeier et al. 2000), as we cover a different portion of the redshift--luminosity plane. Our purpose is to study cosmic source populations at fluxes where the X-ray flux is high enough to provide X-ray spectral information in higher sensitivity follow-up observations. This would allow the determination of the distribution of absorbing columns in the sources making the hard XRB, providing strong constraints on AGN synthesis models for the XRB (e.g. Comastri et al. 1995). ", "conclusions": "" }, "0011/gr-qc0011109_arXiv.txt": { "abstract": "The development of both ground- and space-based gravitational wave detectors provides new opportunities to observe the radiation from binaries containing neutron stars and black holes. Numerical simulations in 3-D are essential for calculating the coalescence waveforms, and comprise some of the most challenging problems in astrophysics today. This article briefly reviews the current status of efforts to calculate black hole and neutron star coalescences, and highlights challenges for the future. ", "introduction": "Binaries containing black holes (BHs) or neutron stars (NSs) are among the most important and interesting sources for the gravitational wave detectors expected to start operation in the early part of the $21^{\\rm st}$ century. These binaries spiral together due to the emission of gravitational radiation, leading to the final collision and coalescence of their components. The NS/NS, NS/BH, and stellar BH/BH binaries are target sources for ground-based interferometers such as LIGO, VIRGO, GEO600, and TAMA, while the space-based LISA is expected to be most sensitive to massive BH binaries (see the articles by Weiss and Kalogera in these Proceedings). Astrophysically, the data from these detectors can provide important insights into NS properties; the equation of state of matter at nuclear densities; models of gamma-ray bursts; active galactic nuclei and quasars; and the strong-field regime of gravity, including unambiguous detection of BH formation (Schutz 1997; Thorne 1998). Calculations of the gravitational waves from such binaries focus on three physical regimes. During the {\\em inspiral phase}, the components are widely separated and can be treated as point particles, allowing analytic calculations using post-Newtonian (PN) expansions. Templates constructed from the resulting waveforms are expected to form a key component of data analysis and source identification (Flanagan \\& Hughes 1998). The {\\em coalescence phase} begins when the compact objects are close enough to suffer tidal deformation, and continues as a combination of relativistic and hydrodynamic effects drives the stars together on more rapid timescales. Calculation of the waveforms produced during this stage requires 3-D numerical simulations that solve the full Einstein equations and, for a NS, general relativistic hydrodynamics; in some cases, the formation of a BH must also be modeled. Finally, the {\\em ringdown phase} encompasses the late-time behavior of the merger product or remnant, radiating in its normal modes (Thorne 1998). Numerical simulation of the coalescence phase using full general relativity is one of the most challenging problems in astrophysics today. This article surveys the current status of efforts to calculate coalescence waveforms for NS/NS, NS/BH, and BH/BH binaries. ", "conclusions": "" }, "0011/astro-ph0011187_arXiv.txt": { "abstract": "We use hydrodynamical $N$-body simulations to study the kinetic Sunyaev--Zel'dovich effect. We construct sets of maps, one square degree in size, in three different cosmological models. We confirm earlier calculations that on the scales studied the kinetic effect is much smaller than the thermal (except close to the thermal null point), with an rms dispersion smaller by about a factor five in the Rayleigh--Jeans region. We study the redshift dependence of the rms distortion and the pixel distribution at the present epoch. We compute the angular power spectra of the maps, including their redshift dependence, and compare them with the thermal Sunyaev--Zel'dovich effect and with the expected cosmic microwave background anisotropy spectrum as well as with determinations by other authors. We correlate the kinetic effect with the thermal effect both pixel-by-pixel and for identified thermal sources in the maps to assess the extent to which the kinetic effect is enhanced in locations of strong thermal signal. ", "introduction": "The Sunyaev--Zel'dovich (SZ) effect (Sunyaev \\& Zel'dovich 1972, 1980; for reviews see Rephaeli 1995 and Birkinshaw 1999) is the change in energy experienced by cosmic microwave background photons when they scatter from intervening gas, especially that in galaxy clusters. The dominant version from clusters is the thermal SZ effect, the gain in energy acquired from the thermal motion of the gas which is commonly at a temperature of tens of millions of degrees in clusters. The kinetic SZ effect is the Doppler shift arising from the bulk motion of the gas. The thermal effect has been quite well studied theoretically, and recently has become a burgeoning area of observational activity with the construction of two-dimensional maps of clusters becoming commonplace (Jones et al.~1993; Myers et al.~1997; Carlstrom et al.~2000). By contrast, less attention has been given to the kinetic effect. It presents a considerable observational challenge, because it is predicted to be much smaller than the thermal effect on the angular scales explored so far, and also because unlike the thermal effect it possesses no characteristic spectral signature allowing it to be distinguished from primary cosmic microwave background anisotropies. It is also more difficult to make semi-analytic calculations. The thermal effect can be obtained fairly directly via the Press--Schechter (1974) approach; various aspects including number counts, the global CMB distortion, and its impact as secondary anisotropy for CMB measurements at small scales have been extensively studied (Cole \\& Kaiser~1988; Colafrancesco et al.~1994, 1997; Bartlett \\& Silk~1994; Barbosa et al.~1996; Eke, Cole \\& Frenk 1996; Aghanim et al.~1997; Komatsu \\& Kitayama 1999; Atrio-Barandela \\& M\\\"ucket 1999; Molnar \\& Birkinshaw 2000; Hern\\'andez-Monteagudo, Atrio-Barandela \\& M\\\"ucket 2000). Predictions for the kinetic effect require the simultaneous estimation of both mass and peculiar velocity, and the signal is much less dominated by the gas which happens to be in massive halos. A detailed analytical model has recently been constructed by Valageas, Balbi \\& Silk (2000) [see also Benson et al.~2000], which computes the effect including inhomogeneities in reionization; this supersedes earlier modelling by Aghanim et al.~(1997) where the peculiar velocities were drawn from a gaussian distribution of fixed width. We recently used large-scale hydrodynamical $N$-body simulations to make simulated maps of the thermal effect and analyzed their properties (da Silva et al.~2000; for similar recent work see Refregier et al.~2000a; Seljak, Burwell \\& Pen 2000; Springel, White \\& Hernquist~2000). In this paper, we make a detailed analysis of maps of the kinetic effect, made using the same technique. Recently Springel et al.~(2000) used similar simulations to study the angular power spectrum of the kinetic effect in the $\\Lambda$CDM cosmology, and investigated the influence of non-gravitational heating upon it. In this paper we study three different cosmologies, and investigate a wide range of properties of the resulting maps including the relation between thermal and kinetic distortion. ", "conclusions": "We have used hydrodynamical simulations to study a number of aspects of the kinetic SZ effect, including its dependence on cosmological parameters. We have studied the redshift dependence, pixel histograms, angular power spectra, and the correlation between the kinetic and thermal effects. We have confirmed that the kinetic effect has a dispersion more than a factor five below the thermal (in the Rayleigh--Jeans region), leading to an angular power spectrum a factor of typically twenty-five lower in the low-density cosmologies. For critical density we have found a smaller difference, but with much greater statistical uncertainty. For the thermal effect the angular power spectrum is mostly generated at redshifts below one, while for the kinetic effect a significant amount of the power has origin above redshift two. The correlation of the kinetic effect with the thermal confirms and quantifies the expected enhancement of the kinetic effect in regions with a strong thermal signal. Our simulations are of ideal size to study the SZ effect on scales between one and several arcminutes, which is currently a resolution attracting great interest. Our results complement perfectly the small-scale simulation work of Bruscoli et al.~(1999) and Gnedin \\& Jaffe (2000), and recent semi-analytic work including that of Benson et al.~(2000) and Valageas et al.~(2000). The theory of these secondary anisotropies is now becoming highly developed; the key challenges for the kinetic effect lie very much on the observational side." }, "0011/astro-ph0011378_arXiv.txt": { "abstract": "We show how estimates of parameters characterizing inflation-based theories of structure formation localized over the past year when large scale structure (LSS) information from galaxy and cluster surveys was combined with the rapidly developing cosmic microwave background (CMB) data, especially from the recent Boomerang and Maxima balloon experiments. All current CMB data plus a relatively weak prior probability on the Hubble constant, age and LSS points to little mean curvature ($\\Omega_{tot} = 1.08\\pm 0.06$) and nearly scale invariant initial fluctuations ($n_s =1.03 \\pm 0.08$), both predictions of (non-baroque) inflation theory. We emphasize the role that degeneracy among parameters in the $L_{pk} = 212\\pm 7$ position of the (first acoustic) peak plays in defining the $\\Omega_{tot}$ range upon marginalization over other variables. Though the CDM density is in the expected range ($\\Omega_{cdm}{\\rm h}^2=0.17 \\pm 0.02$), the baryon density $ \\Omega_b {\\rm h}^2 = 0.030\\pm 0.005$ is somewhat above the independent $0.019\\pm 0.002$ nucleosynthesis estimates. CMB+LSS gives independent evidence for dark energy ($\\Omega_\\Lambda = 0.66 \\pm 0.06$) at the same level as from supernova (SN1) observations, with a phenomenological quintessence equation of state limited by SN1+CMB+LSS to $w_Q < -0.7$ cf. the $w_Q$=$-1$ cosmological constant case. ", "introduction": "\\noindent {\\bf Experiments and Bandpowers:} Anisotropies at the $30 \\mu K$ level at low multipoles revealed by COBE in 1992 were augmented at higher $\\ell$ in some 19 other experiments, some with a comparable number of resolution elements to the 600 or so for COBE, most with many fewer. A list of these experiments to April 1999 with associated bandpowers is given in Bond, Jaffe and Knox (2000 [BJK00]). The anisotropy picture dramatically improved this past year, as results were announced first in summer 99 from the ground-based TOCO experiment in Chile (Miller \\et 2000), then in November 99 from Boomerang-NA, the North American test flight (Mauskopf et 1999). These two additions improved peak localization and gave evidence for $\\Omega_{tot}\\sim 1$. Then in April 2000, results from the first CMB long duration balloon (LDB) flight, were announced (de Bernardis \\et 2000), followed in May 2000 by results from the night flight of Maxima (Hanany \\et 2000). Boomerang's best resolution was $10^\\prime$, about 40 times better than that of COBE, with tens of thousands of resolution elements. Maxima had a similar resolution but covered an order of magnitude less sky. Fig.~1 shows the 150A GHz Boomerang-LDB map and the Wiener-filtered Maxima-1, to scale. The de Bernardis \\et (2000) maps at 90 and 220 GHz show the same spatial features as this 150 GHz one, with the overall intensities falling precisely on the CMB blackbody curve. The Toco, Boomerang and Maxima experiments are described elsewhere in these proceedings. They were designed to reveal the {\\it primary} anisotropies of the CMB, those which can be calculated using linear perturbation theory. Fig.~1 shows the temperature power spectra for Boomerang, Maxima and prior-CMB data (Boomerang-NA+TOCO+April 99) are in good agreement. Sketching the impact of these new results on cosmic parameter estimation (Lange \\et 2000 [Let00], Jaffe \\et 2000 [Jet00]) is the goal of this paper. Space constraints preclude adequate referencing here, but these are given in the Boomerang (Let00) and Maxima+Boomerang (Jet00) parameter estimation papers (see also Bond 1996, [B96], for other references). \\begin{figure} \\vspace{-15pt} \\centerline{\\hspace{30pt}\\epsfig{file=figCLalldata.ps,height=4.5in}} \\vspace{-7pt} \\centerline{\\epsfig{file=figmap.ps,height=3.3in}} \\vspace{-12pt} \\caption{\\small The top figure shows ${\\cal C}_\\ell$ grouped in bandpowers for Boomerang-LDB (crosses), Maxima-I (triangles) and prior-CMB experiments (TOCO+Boomerang-NA+\"April 99\", squares). The lower panel contrasts the optimally-combined power spectra for Boomerang+Maxima+DMR (squares) with that for Boomerang+Maxima+prior-CMB (circles), showing the prior experiments do not move ${\\cal C}_\\ell$ very much. Best-fit models for arbitrary $\\Omega_{tot}$ and for $\\Omega_{tot}$=1 are shown in both panels. The Boomerang 150A GHz map (\\ie for one of 16 bolometers) and the multifrequency Wiener-filtered Maxima-I map, its 124 square degrees drawn to scale, are shown in the bottom figure. Only the 440 square degrees within the central rectangle of the entire 1800 square degrees covered by Boomerang were used in the analysis. \\normalsize } \\label{fig:CLdatmap} \\end{figure} We are only at the beginning of the high precision CMB era for primary anisotropies heralded by the arrival of Boomerang and Maxima, with interferometers taking data (VSA, CBI, DASI), the single dish ACBAR about to, and new LDBs to fly in the next few years (Arkeops, Tophat, Beast/Boost), as well as Boomerang-2001 and the neo-Maxima Maxipol, both concentrating on polarization. In April 2001, NASA's HEMT-based MAP satellite will launch, with $12^\\prime$ resolution, and in 2007, ESA's bolometer+HEMT-based Planck satellite is scheduled for launch, with $5^\\prime$ resolution. \\vskip 10pt \\noindent {\\bf The CMB Analysis Pipeline:} Analyzing Boomerang and other experiments involves a pipeline that takes (1) the timestream in each of the bolometer channels coming from the balloon plus information on where it is pointing and turns it into (2) spatial maps for each frequency characterized by average temperature fluctuation values in each pixel (Fig.~1) and a pixel-pixel correlation matrix characterizing the noise, from which various statistical quantities are derived, in particular (3) the temperature power spectrum as a function of multipole (Fig.~1), grouped into bands, and two band-band error matrices which together determine the full likelihood distribution of the bandpowers (Bond, Jaffe \\& Knox 1998 [BJK98], BJK00). Fundamental to the first step is the extraction of the sky signal from the noise, using the only information we have, the pointing matrix mapping a bit in time onto a pixel position on the sky. To compare the data with millions of cosmological models, as we wish to do here, the radical compression step from 2 to 3 is essential, and hinges upon an accurate representation of the likelihood surface. There is generally another step in between (2) and (3), namely separating the multifrequency spatial maps into the physical components on the sky: the primary CMB, the thermal and kinematic Sunyaev-Zeldovich effects, the dust, synchrotron and bremsstrahlung Galactic signals, the extragalactic radio and submillimetre sources. The strong agreement among the Boomerang maps indicates that to first order we can ignore this step, but it has to be taken into account as the precision increases. The Fig.~1 map is consistent with a Gaussian distribution, thus fully characterized by just the power spectrum. Higher order (concentration) statistics (3,4-point functions, \\etc) tell us of non-Gaussian aspects, necessarily expected from the Galactic foreground and extragalactic source signals, but possible even in the early Universe fluctuations. For example, though non-Gaussianity occurs only in the more baroque inflation models of quantum noise, it is a necessary outcome of defect-driven models of structure formation. (Peaks compatible with Fig.~1 do not appear in non-baroque defect models, which now appear unlikely.) Though great strides have been made in the analysis of Boomerang and Maxima, there is intense needed effort worldwide now to develop new fast algorithms to deal with the looming megapixel datasets of LDBs and the satellites (\\eg Bond \\et 1999, Szapudi \\et 2000). ", "conclusions": "" }, "0011/astro-ph0011470_arXiv.txt": { "abstract": "New results from the search for H$_2$ absorption in the damped Ly$\\alpha$ galaxy at redshift $z = 3.4$ toward QSO 0000--2620 ($z_{\\rm em} = 4.1$) are reported. The high-resolution ($\\lambda/\\Delta\\lambda = 48,000$) spectra of Q0000--2620 were obtained using the Ultraviolet - Visual Echelle Spectrograph (UVES) on the 8.2m {\\itshape ESO} Kueyen telescope. The ortho-H$_2$ column density is found to be $N(J=1) = (5.55 \\pm 1.35)\\times10^{13}$ cm$^{-2}$ ($2\\sigma$ C.L.). The combination of $N(J=1)$ with the limits available for other low rotational levels restricts the excitation temperature $T_{\\rm ex}$ in the range $(290 - 540)$~K. This gives the total H$_2$ column density of $N({\\rm H}_2) = (8.75 \\pm 1.25)\\times10^{13}$ cm$^{-2}$ and the corresponding fraction of hydrogen atoms bound in molecules of $f({\\rm H}_2) = (6.8 \\pm 2.0)\\times10^{-8}$. ", "introduction": "It has long been recognized that H$_2$ (and HD) molecules play a central role in the formation of gas condensations in the post-recombination era since they provide the cooling necessary for the collapse on all scales of the first objects. In the primordial gas at redshift $z \\la 50$, the fractional abundance of H$_2$ is calculated to be $f({\\rm H}_2) = 10^{-5} - 10^{-6}$ (e.g. \\cite{ref1.1}). The H$_2$ absorption lines from the Lyman and Werner bands may be observable at lower redshift $z \\sim 2 - 4$ in QSO absorption systems with high neutral hydrogen column densities, $N({\\rm HI}) = 10^{21} - 10^{22}$ cm$^{-2}$ (so-called Damped Lyman $\\alpha$ systems, DLA). It has been suggested that these systems are most closely related to the progenitors of normal galaxies~\\cite{ref1.2}. So far, absorption from H$_2$ has been detected in a few DLA systems which show, in general, a small amount of molecular gas~\\cite{ref1.2a}. In contrast to DLAs, observations in our Galaxy reveal the presence of H$_2$ in nearly all lines of sight in the disk and halo~\\cite{ref1.3}. It was also found that the ISM diffuse clouds with the neutral hydrogen column densities $N({\\rm HI}) \\ga 3\\times10^{20}$ cm$^{-2}$ show $f({\\rm H}_2) \\ga 10^{-5}$. \\begin{figure} \\vspace{-1.5cm} \\centering \\includegraphics[width=1.0\\textwidth]{lev_fig1.eps} \\vspace{-1.0cm} \\caption[]{ Best fit to H$_2$ lines associated with the $z = 3.4$ DLA system toward Q0000--2620. ({\\bf a}, {\\bf b}, {\\bf c}, {\\bf d}) Simultaneous fit to the Lyman L(4-0)R(1), L(1-0)R(2) and Werner W(2-0)Q(1), W(2-0)R(0) lines with $N(J=0) = 8.4\\times10^{12}$ cm$^{-2}$ ($1\\sigma$ upper limit), $N(J=1) = (5.5 \\pm 0.7)\\times10^{13}$ cm$^{-2}$, $N(J=2) = 1.2\\times10^{13}$ cm$^{-2}$ ($1\\sigma$ upper limit), and $b \\simeq 10.07$ km~s$^{-1}$. The velocities shown are related to $z = 3.390127$. ({\\bf e}) Confidence regions for the ortho-H$_2$ column density calculated from the simultaneous fit of the L(4-0)R(1) + W(2-0)Q(1) lines and all low-ion lines located redward the Ly$\\alpha$ emission. ({\\bf f}, {\\bf g}) The corresponding deviations of the local continuum level. ({\\bf h}) The corresponding Doppler parameter variations. The grey areas restrict the $\\Delta C / C$ and $b$ values at $1\\sigma$ level in accordance with panel {\\bf e} } \\label{eps1} \\end{figure} Absorption from H$_2$ in the $z = 3.4$ DLA system toward Q0000--2620 was not detected in the 1 \\AA\\, resolution spectrum obtained with the Multiple Mirror Telescope (MMT) and only an upper limit of $f({\\rm H}_2) < 3\\times10^{-6}$ has been reported previously~\\cite{ref1.4}. New observations with the UVES/VLT, having approximately 11 times higher resolution, revealed H$_2$ absorption in this DLA. Exactly at the expected position of the L(4-0)R(1) line where the MMT gave the limit $W_{\\rm rest} < 114$ m\\AA\\, (3$\\sigma$), an absorption line with the equivalent width $W_{\\rm rest} \\simeq 6$ m\\AA\\, was detected~\\cite{ref1.5}. The chance identification of this line in the Ly$\\alpha$ forest at $z \\sim 3$ has a probability $\\la 10^{-3}$ (cf.~\\cite{ref1.5a}). Below we report on new identifications of H$_2$ lines in this DLA system and give improved values of molecular hydrogen column density and excitation temperature. ", "conclusions": "" }, "0011/astro-ph0011536_arXiv.txt": { "abstract": "Interferometric techniques are at the forefront of modern astronomical instrumentation. A new generation of instruments are either operating or nearing completion, including arrays of small telescopes as well as the ``big guns'' (VLTI and Keck). A number of space interferometers for the detection of extra-solar planets are also being planned. I will review the current state of play and describe the latest developments in the field. ", "introduction": "It is somewhat disappointing that a meeting on ``Galaxies and their Constituents at the Highest Angular Resolutions'' contains so few results from optical inter\\-ferometry\\footnote{In this review, the term ``optical'' is also meant to include infrared wavelengths.}. To some extent, this reflects the interests of those on the Scientific Organising Committee, but it must also be taken as a sign that optical interferometry has yet to deliver on its promises. Why is this? One reason is surely the difficultly of the technology and its vulnerability to Hofstadter's Law (1980): \\begin{quotation} {\\em Hofstadter's Law} --- It always takes longer than you expect, even when you take into account Hofstadter's Law. \\end{quotation} But progress is being made, and there is a large number of interferometry projects in various stages of development and operation. The following is the current list of operating interferometers, listed in the order in which they obtained first fringes: \\begin{itemize} \\itemsep=0pt \\item GI2T/REGAIN (Grand Interf\\'erom\\`etre \\`a 2 T\\'elescopes). \\item SUSI (Sydney University Stellar Interferometer) \\item COAST (Cambridge Optical Aperture Synthesis Telescope) \\item ISI (Infrared Spatial Interferometer) \\item FLUOR (Fiber Linked Unit for Optical Recombination) \\item IOTA (Infrared-Optical Telescope Array) \\item NPOI (Navy Prototype Optical Interferometer) \\item PTI (Palomar Testbed Interferometer) \\item CHARA (Center for High Angular Resolution Astronomy) \\end{itemize} The following are under construction (in alphabetical order, since I am not game to predict which will be first to get fringes!): \\begin{itemize} \\itemsep=0pt \\item Keck Interferometer \\item LBT (Large Binocular Telescope) \\item MIRA-II (Mitaka IR Array) \\item VLTI (Very Large Telescope Interferometer) \\end{itemize} Finally, there are planned space missions, including SIM, TPF and DARWIN\\@. More details of all these projects can be found via links from the Web-based ``Optical Long Baseline Interferometry News,'' currently maintained by Peter Lawson\\footnote{\\tt http://huey.jpl.nasa.gov/olbin/}. Here, I simply want to stress the number and range of the projects, and to point out that many of them are producing scientific results. The proceedings of recent conferences (Unwin \\& Stachnik 1999; Lena \\& Quirrenbach 2000) testify to the vigorous activity in this field. In Section~\\ref{sec.recent}, I highlight some areas in which recent technological developments have been made. First, however, it is appropriate to make some general remarks about optical interferometry. ", "conclusions": "" }, "0011/astro-ph0011299_arXiv.txt": { "abstract": "We reanalyse the UV/optical spectrum and optical broad-band data of the eclipsing binary HV 2274 in the LMC, and derive its distance following the method given by Guinan et al.~(1998a,b) of fitting theoretical spectra to the stars' UV/optical spectrum plus optical photometry. We describe the method in detail, pointing out the various assumptions that have to be made; moreover, we discuss the systematic effects of using different sets of model atmospheres and different sets of optical photometric data. It turns out that different selections of the photometric data, the set of model atmospheres and the constraints on the value of the ratio of selective to total extinction in the $V$-band, result in a 25\\% range in distances (although some of these models have a large ${\\chi}^2$). For our best choice of these quantities the derived value for the reddening to HV 2274 is $E(B-V)$ = 0.103 $\\pm$ 0.007, and the de-reddened distance modulus is DM = 18.46 $\\pm$ 0.06; the DM to the center of the LMC is found to be 18.42 $\\pm$ 0.07. This is significantly larger than the DM of 18.30 $\\pm$ 0.07 derived by Guinan et al. (1998a). ", "introduction": "The distance to the Large Magellanic Cloud (LMC) is a fundamental step in the cosmological distance ladder, since the extragalactic distance scale is usually determined with respect to the LMC distance. In fact, both the {\\sc hst} $H_0$ Key Project (Kennicutt et al. 1995, Freedman et al. 1999) and the Supernovae Calibration Team (Saha et al. 1999) fix the zero-point of the cosmological distance scale assuming a de-reddened LMC distance modulus (DM) of 18.50; in the case of the {\\sc hst} $H_0$ Key Project the adopted uncertainty of $\\pm$ 0.13 on the LMC distance modulus represents the largest contribution to their systematic error budget. In recent years various methods have yielded LMC distance moduli showing a remarkable spread, ranging from DM = 18.07 (Udalski et al. 1998a) to DM = 18.70 (Feast \\& Catchpole 1997) -- see for example the review by Gibson (1999) or Feast (2001). This uncertainty alone on the LMC distance causes an indetermination by $\\sim$20\\% on the value of the Hubble constant. A method that, at present, seems to support the short distance scale, involves the analysis of the light-curve, radial velocity curve and UV/optical spectrum of the detached eclipsing binary HV 2274 in the LMC. It is considered (see, e.g., Gibson 1999) to be one of the most promising techniques to derive a precise distance to the LMC, and it is based on a very elegant idea. From the analysis of the radial velocity and light-curve one obtains very accurate values for the masses and radii of the two binary components, as well as for the ratio of the effective temperatures. Fitting the UV/optical spectrum with model atmospheres one obtains the reddening, effective temperature and distance. This method, as already stressed, has been put forward as {\\sl the} way to obtain a very accurate DM to the LMC, in particular when more systems will be analysed. Guinan et al. (1998a; hereafter G98a) found DM = 18.30 $\\pm$ 0.07 (the derived distance modulus to HV 2274 was 18.35 $\\pm$ 0.07. A geometric correction has been then applied to obtain the distance to the center of the LMC) after applying this technique. More recently Nelson et al.~(2000) corrected the G98a value after re-determining the reddening toward HV 2274; they obtained DM = 18.40 $\\pm$ 0.07, a value only marginally in agreement with the long distance. However, they did not apply the method by G98a, deriving the distance correction only in an indirect way. Because of the relevance of the method employed by G98a for deriving the LMC distance and in light of the recent claims by Nelson et al.~(2000) for a longer LMC distance, we want in the present paper to re-analyse the fitting procedure of the UV/optical spectrum, and carefully point out the uncertainties of this method. Unfortunately, in their 4-page {\\sl Letter}, G98a could not present all the intricate details that {\\sl are} involved in the application of this method, but which should be pointed out to the scientific community in order to judge the strengths and weaknesses of this technique (also see Feast 2001). We will also study the sensitivity of the derived distance to both the set of model atmospheres and of optical photometric data employed in the fitting procedure. In Sect.~2 we present an historical overview of observations and studies related to the HV 2274 distance. In Sect.~3 and 4, we discuss, respectively, the observational data and the method employed for the distance determination. Results are presented in Sect.~5, while a discussion follows in Sect.~6. ", "conclusions": "We will now discuss the results obtained in the previous section, with particular emphasis on the consistency of the derived parameters for the reference model. \\subsection{The distance to HV 2274} As previously explained, our reference model is model 5, which has the maximum wavelength coverage -- from the UV to the near infrared -- of the HV2274 spectrum, and makes use of the set of theoretical spectra which provides the lower $\\chi^2$ value. In Fig.~1 and 2 we show, for model 5, the fit of the theoretical spectra to the FOS one, and the resulting UV and optical normalized extinction curve. The final value for $\\left( \\frac{r_{\\rm A}}{d}\\right)^2$ is that of model 5. The final error estimate comes from the formal error in the model 5 result, added in quadrature to the squared differences of the parameter values of model 5 with those of models 7-16, and the additional error due to the error in $r_{\\rm A}$. The result is $\\left( \\frac{r_{\\rm A}}{d}\\right)^2$ = (2.05 $\\pm$ 0.12) 10$^{-23}$. This corresponds to a linear distance of 49.10 $\\pm$ 1.44 kpc, or a true distance modulus of 18.46 $\\pm$ 0.06. \\subsection{The distance to the center of the LMC} To obtain the distance to the center of the LMC, the location of the binary with respect to the LMC center has to be taken into account. The geometry of the LMC can be described by an inclined disk, and one therefore has to consider the distance from the plane of that disk to the plane of the sky through the center of the LMC at the position of the binary, and the fact that the binary may be behind or in front of the plane of the disk. G98a assume that HV 2274 is 1.1 kpc behind the center of the LMC, based on the parameters of Schmidt-Kaler \\& Gochermann (1992). Various authors have described the geometry of the LMC with a thin disk and derived the position angle ($\\theta$) of the line-of-nodes and inclination angle ($i$). The relevant coordinate transformation to go from observed right ascension and declination to a rectangular coordinate system in the plane of the sky, and to a similar one rotated by $\\theta$ and $i$ are given in Weinberg \\& Nikolaev (2000). Table~6 gives estimates of the distance of the LMC plane to the center of the LMC for various accurate estimates of $\\theta$ and $i$, using the coordinate system by Weinberg \\& Nikolaev (2000). Because of the different orientations and definition of positive inclination the values in the table may differ by 90\\degr\\ with respect to the values quoted in the original reference. The mean of these 5 determinations is 0.88 kpc; the average of the highest and lowest value is 0.89 kpc and the median is 0.82 kpc. The adopted difference in distance is 0.9 $\\pm$ 0.3 kpc\\footnote{As an aside we did the same for SN 1987A. The shortest distance (SN 1987A - LMC Center) is $-0.64$ kpc [for the parameters of Schmidt-Kaler \\& Gochermann (1992)], and the largest is $-0.15$ kpc [for the parameters of Groenewegen (2000)]. Based on all 5 determinations, the LMC plane at the location of SN1987A is 0.4 $\\pm$ 0.2 kpc in front of the LMC center.}. The error does not come from the internal errors of each of the determinations, but from the spread among the values itself, and the difference between the lowest and highest value has been assumed to correspond to 3$\\sigma$. It is unknown if HV 2274 is in front or behind the LMC plane. The vertical scale height of the LMC disk is small however, between 100 and 300 pc (see the discussion in Groenewegen, 2000). The largest value is taken here, and added to the error mentioned above, to give the final result that HV 2274 is located 0.9 $\\pm$ 0.5 kpc behind the LMC center. Considering a DM to HV 2274 of 18.46 $\\pm$ 0.06, the DM to the LMC center 18.42 $\\pm$ 0.07. \\begin{table} \\caption[]{The distance (HV 2274 $-$ LMC center) } \\begin{tabular}{cccccccc} \\hline $\\theta$ & $i$ & Reference & $\\Delta$ \\\\ (\\degr) & (\\degr) & & (kpc) \\\\ \\hline 258 & 38 & Schmidt-Kaler \\& Gochermann (1992) & 1.30 \\\\ 258 & 33 & Feitzinger et al. (1977) & 1.08 \\\\ 261 & 25 & Weinberg \\& Nikolaev (2000) & 0.82 \\\\ 296 & 18 & Groenewegen (2000) & 0.75 \\\\ 232 & 29 & Martin et al. (1979) & 0.47 \\\\ \\hline \\end{tabular} \\end{table} \\subsection{The extinction curve towards HV 2274} From model 5 we get $x_0 = 4.67 \\pm 0.04$, $\\gamma = 1.27 \\pm 016$, $c_3 = 2.03 \\pm 0.58$, $c_4 = 0.66 \\pm 0.10$ and $R = 3.06 \\pm 0.14$ (where the errors are the internal errors due to the fitting procedure only); from Eqs.~(6) and (7) one obtains $c_1 = -0.13 \\pm 0.21$ and $c_2 = 0.72 \\pm 0.07$. These values are consistent with the corresponding quantitites recently determined by Misselt et al.~(1999) for their 'LMC-Average Sample' of stars. Figure 2 shows the resulting normalised UV and optical extinction curve. \\subsection{The reddening towards HV 2274} The value of $E(B-V)$ obtained from model 5 is $E(B-V)$ = 0.103 $\\pm$ 0.007. It is important to notice that this value is different from the one derived by G98b ($E(B-V)$ = 0.120 $\\pm$ 0.009) and from the determination by UPW98 ($E(B-V)$ = 0.149 $\\pm$ 0.015) who used colour-colour $(U-B)-(B-V)$ relationships. Since we are using the $(B-V)$ (and $(V-I)$) photometric data by UPW98 as constraint for the spectrum fit we investigated the possibility that the reddening derived from the spectrum fit is inconsistent with HV2274 broad band photometry. We followed the same procedure by UPW98 that is, to consider a local standard $(U-B)-(B-V)$ sequence of stars with the same spectral type as HV 2274, and derive the reddening from the displacement of the position of HV 2274 with respect to the standard sequence, for an assumed $E(U-B)/E(B-V)$ ratio. We used the same standard colour-colour sequence employed by UPW98 (for B-stars of luminosity class III), and a ratio $E(U-B)/E(B-V)$ derived from our adopted extinction law using the value of $R$ derived from the fit (which is by the way very close to the standard value of 3.1). By computing appropriate stellar models (using the same input physics and colour transformations as in Salaris \\& Weiss 1998 and a scaled-solar metal distribution) for solar metallicity and the metallicity derived from model 5, with masses around 12$M_{\\odot}$, we verified that the location of B stars on the $(U-B)-(B-V)$ plane does not vary in this metallicity range. As for the $(B-V)$ colour of HV 2274 we used the value by UPW98, while for $(U-B)$ we derived the value from the FOS spectrum (see Section 4.2), since it covers the wavelength region spanned by this colour index. The $(U-B)$ colour derived from the spectrum is different from the value observed by UPW98. The FOS spectrum provides $(U-B)$ = $-$0.836 $\\pm$ 0.006, while UPW98 measured $(U-B)$ = $-$0.905 $\\pm$ 0.04. Nelson et al.~(2000) measured $(U-B)$ = $-0.793$ $\\pm$ 0.031 which is closer (but still inconsistent at the 1$\\sigma$ level) to the value derived from the FOS spectrum. From the colour-colour diagram we get $E(B-V)$ = 0.115 $\\pm$ 0.015 which is consistent, within 1$\\sigma$, with the value derived from model 5. \\subsection{The metallicity of HV 2274} The value of [m/H] from the reference model, taking the formal error from the fit plus the external errors from models 7-16 is $-0.38 \\pm 0.12$. This is relative to the adopted iron abundance of 7.67 in the Kurucz models. This implies that the metallicity relative to the currently favoured solar iron abundance of 7.51 (Grevesse \\& Noels 1993) is [Fe/H] = $-0.22 \\pm 0.12$. Model 6 computed using Butler atmospheres has [m/H] = $-0.13$ relative to a solar abundance of 7.46, [Fe/H] = $-0.18$ relative to the solar abundance of Grevesse \\& Noels. These values are in good agreement with each other, but significantly more metal rich than the value of $-0.42 \\pm 0.07$ to $-0.45 \\pm 0.06$ found by G98a,b. Observationally there are few direct iron abundance determinations for hot and young stars in the LMC. Haser et al. (1998) derived a metallicity of $-0.3$ and $-0.1$ dex for an O3{\\sc iii} and O4{\\sc i} star in the LMC. Korn et al. (2000) determined the iron abundance in 5 non-supergiant B-stars, the average value being $-0.42 \\pm 0.15$. Both the values found by G98a,b and ours are consistent with these determinations. \\subsection{The \\M\\ data} As mentioned in Sect.~3 HV 2274 has been detected in the \\M\\ $JHK_{\\rm s}$ infrared survey. The observation date is JD = 2451111.689. From the ephemeris in Watson et al. (1992) a primary eclipse is predicted at JD = 2451111.697 $\\pm$ 0.006, taking into account the error in the time determination of the reference primary eclipse and the period. The shift of 0.008 $\\pm$ 0.006 days, or 0.0014 in phase is negligible, also in light of the fact that phase shifts of up to 0.05 occur due to apsidal rotation (Watson et al. 1992). In the optical the magnitude difference between out-of-eclipse and primary eclipse are 0.72, 0.71 and 0.70 mag in $BVI$, respectively (Watson et al. 1992). Naively, one might therefore expect a magnitude difference of about 0.69 in $J$. The difference between the observed \\M\\ $J$ and the model predictions is (15.152 $\\pm$ 0.057) $-$ (14.503 $\\pm$ 0.049) = 0.65 $\\pm$ 0.08, where the error in the model comes from summing up all the error terms from models 6-17. This is in good agreement with the value extrapolated from the magnitude difference between out-of-eclipse and primary eclipse measured in the optical. However, as emphasised later, out-of-eclipse NIR photometry would be important in further constraining the parameters of the model. \\begin{figure} \\centerline{\\psfig{figure=h2475f2.ps,width=8.5cm}} \\caption[]{UV and optical normalised extinction curve for the parameters of model 5.} \\end{figure} \\subsection{Final remarks} In this paper we have described and analyzed in detail the method used by G98a,b to derive the distance to the eclipsing binary HV 2274 in the LMC. We used various sets of theoretical spectra and broad band photometric data in the fitting procedure outlined in Section 4, and we found that Kurucz ATLAS 9 spectra best reproduce the observed spectrum of HV 2274. The selection of the wavelength range to be covered by the spectrum fit and the constraint on $R$ play also a role in determining the outcome of the fitting procedure. We are now going to comment briefly about this point. In the reference model we fitted HV 2274 FOS data plus $(B-V)$ and $(V-I)$ colours (from UPW98) using ATLAS 9 spectra, constraining the parameters $c_1$ and $c_2$ of the UV extinction law according to Equations (6) and (7) and keeping $R$ as a free parameter. The use of $(V-I)$ is dictated by our desire to use all available information about the spectral energy distribution of HV 2274. The homogeneous $UBVI$ photometry by UPW98 makes it possible to cover all the spectral range from UV to near-infrared. Neglecting $(V-I)$, that means, fitting a smaller wavelength range, induces a decrease by 0.10 mag in the distance modulus derived from the fitting procedure. If we consider the ATLAS 9 models as an accurate reproduction of the 'real' spectra of B stars, this difference can be ascribed to the fact that there is information about the distance contained in the $(V-I)$ colour, and therefore it must be included in the fitting procedure. Conversely, this result could also imply that ATLAS 9 spectra are inconsistent with observations at the longer wavelengths or that broad-band photometry and FOS spectrum are not homogeneously calibrated. In this respect, one should keep in mind that the $(U-B)$ value derived from the FOS spectrum is inconsistent with the $(U-B)$ determined by UPW98 and Nelson et al.~(2000) and also that the $(U-B)$ values derived from the best-fit ATLAS 9 spectrum are systematically different by $\\sim$0.03 mag with respect to the FOS colour (see Table~5). In case of not considering $(V-I)$ for the spectrum fitting, keeping $R$ fixed increases sensibly (0.09 mag) the derived distance modulus with respect to the case of having $R$ determined by the fitting procedure, as shown by comparing the outcome of models 21 and 17. However, it is not clear why one should fix the value of $R$ a priori, since it is known that it is subjected to variations within the LMC and within our galaxy. Finally, the use of the $(B-V)$ by Nelson et al.~(2000) together with the FOS spectrum produces a distance modulus even larger (0.16 mag larger than the case of using UPW98 data). The quality of the fit is only marginally lower, and the reddening compares well with the value ($\\sim$ 0.08) one would derive from the colour-colour diagram considering the Nelson et al.~(2000) $(B-V)$ and the FOS $(U-B)$. A last comment is that it is very unfortunate that the \\M\\ infrared data happens to be taken during eclipse. The difference in predicted $JHK$ between the different models is up to 0.15 mag (see Table~5). Therefore, accurate NIR photometry or even NIR spectroscopy at the 1\\% level is expected to give valuable additional constraints. \\\\ The conclusion is that the use of eclipsing binaries as distance indicators is powerful, but that a photometrically well calibrated data set covering a large wavelength region is essential. Furthermore, the discrepancies among the different theoretical model atmospheres is worrying and need further investigation." }, "0011/astro-ph0011250_arXiv.txt": { "abstract": "We explore the relative importance of the stellar mass density as compared to the inner dark halo, for the observed gas kinematics thoughout the disks of spiral galaxies. We perform hydrodynamical simulations of the gas flow in a sequence of potentials with varying the stellar contribution to the total potential. The stellar portion of the potential was derived empirically from K-band photometry. The output of the simulations -- namely the gas density and the gas velocity field -- are then compared to the observed spiral arm morphology and the H$\\alpha$ gas kinematics. We solve for the best matching spiral pattern speed and draw conclusions on how massive the stellar disk can be at most. For the case of the galaxy NGC 4254 (Messier 99) we demonstrate that the prominent spiral arms of the stellar component would overpredict the non-circular gas motions unless an axisymmetric dark halo component adds significantly in the radial range ${\\rm R_{exp} < R < 3\\, R_{exp}}$. ", "introduction": "In almost all galaxy formation scenarios, non-baryonic dark matter plays an important role. Today's numerical simulations of cosmological structure evolution quite successfully reproduce the observed galaxy distribution in the universe \\cite{kau}. While galaxies form and evolve inside dark halos their physical appearance depends strongly on the local star formation and merging history. At the same time the halos evolve and merge as well. According to the simulations, we expect that the dark matter is important in the inner parts of galaxies \\cite{NFW1},\\cite{NFW2} and that it thus has a considerable influence on the kinematics. These predictions are in contrast to some studies which indicate that galactic stellar disks - at least of barred spiral galaxies - alone dominate the kinematics of the inner regions \\cite{deb}. Apparently this is also the case in our own Milky Way \\cite{ger}. \\\\ Determining individual mass fractions of the luminous and dark matter is not a straightforward task. The rotation curve of a disk galaxy is only sensitive to the total amount of gravitating matter, but does not allow the distinction between the two mass density profiles. For a detailed analysis it is necessary to adopt more refined methods to separate out the different profiles. Previous investigations used for example knowledge of the kinematics of rotating bars \\cite{wei} or the geometry of gravitational lens systems \\cite{mal}. Here we would like to exploit the fact, that the stellar mass in disk galaxies is often organized in spiral arms, i.e.~in coherent non-axisymmetric structures. In most proposed scenarios, the dark matter, however, is collisionless and dominated by random motions. Therefore it is not susceptible to spiral structures. If the stellar mass dominates, the spiral arms, as traced by the near infrared (NIR) light, should induce considerable non-circular motions in the gas, that manifest themselves as velocity ''wiggles'' in observed gas kinematics. Using hydrodynamical gas simulations we are able to predict these velocity wiggles and compare them to the observations. Hence the contribution of the perturbative forces with respect to the total forces can be determined quantitatively and can be used to constrain the stellar disk to halo mass ratio. ", "conclusions": "Although there is quite some scatter in the observed gas kinematics, we find that the velocity jumps, which are apparent in the simulations for the maximum disk case are too large to be in agreement with the measurements. The inner part of the simulated rotation curve ($<$ 0\\farcm3) is dominated by the dynamics of the small bar, which is present at the center of the galaxy. Its pattern speed might be different from the one of the spiral's and thus relate to a mismatch in the inner part of the rotation curve. We conclude that an axisymmetric dark halo is needed to explain the kinematics of the stellar disk. The influence of the stellar disk is submaximal in the sense that we don't find strong enough velocity wiggles in the observed kinematics as would be expected if the stellar disk was the major gravitating source inside the inner few disk scale lengths. How this conclusion might apply to other spiral galaxies will be the upcoming issue of this project. We plan to extend our analysis at first to the 3 other galaxies where we have already now complete data sets. Finally we intend to draw our final conclusions on a basis of a sample consisting of 8 - 10 members. This should be sufficient to determine reliable results about the luminous and dark mass distributions in spiral galaxies." }, "0011/astro-ph0011066_arXiv.txt": { "abstract": "Deep submillimetre surveys have successfully detected distant, star-forming galaxies, enshrouded in vast quantities of dust and which emit most of their energy at far infrared wavelengths. These luminous galaxies are an important constituent of the Universal star-formation history, and any complete model of galaxy evolution must account for their existence. Although these sources have been tentatively identified with very faint and sometimes very red optical counterparts, their poorly constrained redshift distribution has made their interpretation unclear. In particular, it was not understood if these galaxies had been missed in previous surveys, or if they constituted a truly new class of objects, undetectable at other wavelengths. By utilizing a radio selection technique, we have isolated a % sample of 20 sub-mm objects representative of the 850\\mum\\ population brighter than 5\\,mJy with $z$\\cle 3. We show that these galaxies are so heavily dust obscured that they remain essentially 'invisible' to ultraviolet selection. Furthermore, relying on the radio-submillimeter flux density ratio, we estimate their redshift distribution, finding a median of two. These results are inconsistent with the existence of a very high redshift ($z>4$) population of primeval galaxies (L$_{bol} > 10^{12}$\\,h$^{-2}$\\Lsun) contributing substantially to the sub-mm counts. While not a substitute for the thorough followup of blank field sub-mm surveys, our results do shed light on a substantial portion of the luminous sub-mm population with $z$\\cle 3. ", "introduction": "The extragalactic far-infrared background light is believed to be composed of the integrated thermal starlight and non-thermal AGN radiation, reradiated by dust within star-forming galaxies over the entire history of galaxy formation. The energy density of this infrared background is approximately the same as found in the optical suggesting that at least half of the Universal star-formation history remains hidden from optical view (Puget et al.~1996). This diffuse background was first resolved into discrete sources by the Sub-millimetre Common User Bolometer Array (SCUBA -- Holland et al.~1999) on the James Clerk Maxwell Telescope (JCMT) by a number of groups (Smail et al.~1997, Hughes et al.~1998, Barger et al.~1998, Eales et al.~1999). Although a large number of deep SCUBA surveys has led to a better estimate of the 850 micron galaxy surface density, our understanding of the nature of the sub-mm population remains limited. The principal obstacle is obtaining reliable counterparts of these sub-mm sources at other wavelengths, a problem exacerbated by both the coarseness of the JCMT resolution (15 arcsec at 850 microns) and the inherent faintness of suspected optical counterparts (Smail et al.~1999). It is still unclear if the sub-mm selected sources are related to known populations, such as high redshift quasars (Hughes et al.~1997, McMahon et al.~1999) or Lyman-break galaxies (Chapman et al.~2000a), or constitute a truly new class of objects. With the exception of several isolated objects, few reliable identifications have been made (e.g. Ivison et al.~1998, Frayer et al.~1998). Thus the redshift distribution has remained largely unconstrained over a vast range, with the possibility that many sources lay at extreme distances ($z > 5$). One technique, which has shown some promise in identifying sub-mm sources, is radio continuum followup. Because galaxies and the inter-galactic medium are transparent at centimeter wavelengths, radio emission is unhindered by intervening gas and dust. Ubiquitous in local star-forming galaxies, radio emission also correlates very strongly with the far-infrared emission in star-forming galaxies (Helou et al.~1986). Moreover, the high resolution provided by radio interferometers can provide a surrogate for the poor sub-mm angular resolution and astrometric uncertainties. Given the difficulties of obtaining secure optical identifications and spectroscopic redshifts, the radio observations provide another clue to the nature of sub-mm sources. Via the empirically observed far-infrared to radio correlation in local star-forming galaxies, one can use the observed ratio of sub-mm to radio continuum flux density to estimate a redshift. As the k-corrections (corrections based on the redshifted spectral energy distribution - SED) of the radio and sub-mm flux densities are opposite in slope, the ratio of radio to sub-mm flux density is quite sensitive to redshift (Carilli \\& Yun 1999). Barger, Cowie \\& Richards (2000 -- hereafter BCR) first attempted to use a radio selected sample to target a number of optically faint microJansky radio sources with near-infrared magnitude, $K > 20.5$. Using SCUBA to a 3$\\sigma$ RMS limiting flux density of 6\\,mJy at 850 microns, they detected 5 out of 15 radio sources, while in the process demonstrating that none of the optically brighter radio sources ($K < 20.5$) were detected in the sub-mm. The surface density of these few bright radio selected sub-mm sources closely matched that from blank field surveys, indicating a close correspondence between the optically faint radio population and bright sub-mm sources. Other pointed SCUBA studies of known high-$z$ populations such as $z\\sim3$ Lyman-break galaxies (LBGs -- Chapman et al.~2000a) and radio loud AGN (Archibald et al.~2000) have revealed few SCUBA detections, and nowhere near the surface density of blank field sub-mm sources. We have refined the selection criterion to those microJansky radio sources with an optical magnitude, $I > 25$, based on the clear bi-modal break in the optical properties of microJy radio sources (Richards et al.~1999). We have applied this technique to a sample in the region surveyed by Richards (2000) centered on the Hubble Deep Field. We have selected a total of 47 radio sources in our study, 20 previously observed, which meet our criterion. Our followup SCUBA {\\it photometry} observations of 27 radio selected objects demonstrate this to be a highly successful technique. We now detect $\\sim$50\\% of the new 27 object sample above 4.5\\,mJy at 850 microns, with an overall success rate of 20 out of 47 objects observed. Our increased detection success over BCR is likely to be a result of our slightly deeper survey coupled with the stricter selection criterion. We are thus able to uncover bright sub-mm sources using SCUBA at the rate of one source per hour on the JCMT, greater than an order of magnitude more rapid than mapping a random patch of sky. Our new survey represents a sub-mm mapping of a $\\sim$100 arcmin$^2$ effective region in less than 16 hours, sensitive to sources S$_{850}>5$\\,mJy and $z$\\cle 3. ", "conclusions": "The crucial data available to us from our technique are the optical properties and redshift estimates for the sub-mm sources, which we present in Table~1. Our results assume a $\\Lambda=0.0$, $\\Omega=1.0$, H$_0$=65\\,km/s/Mpc cosmology. Redshift estimates can be obtained from the sub-mm/radio flux ratios (Carilli \\& Yun 2000). All of our sub-mm sources fall roughly in the redshift range $z=1-3$ with a median redshift for the sample of $z=1.9$, consistent with previous results from BCR. The sensitivity of our radio survey to star-forming galaxies with radio luminosities fainter than $10^{24}$\\,W/Hz diminishes quickly past $z\\sim3$, and hence biases our sub-mm survey. An independent check on the sub-mm/radio redshift estimates can be obtained through the 450\\mum/850\\mum\\ ratio (e.g. Hughes et al.~1998). Subject to unknown dust temperature, we obtain an estimate of T$_{\\rm d}$/(1+$z$), which we list in Table~1 for T$_{\\rm d}$=45\\,K for consistency with the ultra-luminous infrared galaxy, Arp\\,220. Raising or lowering the adopted dust temperature has the effect of a corresponding systematic raising and lowering of both our redshift estimates (Blain 1999). In order to calculate the density of sub-mm sources on the sky as found in our radio pre-selection survey, we need to determine the effective area covered by our study. This is given simply by the overlap regions between the optical and radio images. However, a further complication arises from the non-uniform sensitivity of the radio image which serves to decrease the visibility area. This issue is discussed in Richards (2000) and we use the same method for determining the source count. Our pre-selected sample has already saturated the bright ($>5$\\,mJy) sub-mm counts (Fig.~1), % and there are not likely to be many additional bright, high redshift sub-mm sources in our survey region. BCR found that 2 additional sub-mm sources without radio counterparts were detected in their survey area, indicating 75\\% of the bright sub-mm sources are typically recovered through such radio pre-selection. Since we performed photometry on the sources, we have no means of estimating this extra population. By accounting for such an additional 25\\% high redshift population our sample is in agreement with analyses of the redshifts for lensed sub-mm sources (Barger et al.~1999, Smail et al.~2000, Blain et al.~1999b). The percentage of sub-mm sources missed by our pre-selection technique will depend field to field on the high-$z$ clustering of sub-mm luminous sources. A lensed sub-mm survey (Smail et al.~2000) detects a similarly large fraction of their bright sub-mm sources in the radio. Deeper blank field sub-mm surveys (e.g. Eales et al.~2000 -- S$_{850\\mu m}$\\cge 3\\,mJy) detect $\\sim$1/3 of their sources in the radio. This is roughly in agreement with our results which use deeper radio maps and brighter sub-mm limits. While not a substitute for the thorough (and difficult) followup of blank field sub-mm surveys, our results do shed light on a substantial portion of the luminous sub-mm population with $z$\\cle 3. At the faint end of the counts, our pre-selected sources appear to fall short of full source counts recovery, and the interesting question becomes what is the nature of the 850-micron sources at flux densities $<$5\\,mJy? Although we are only sensitive to sources with S$_{850}$\\cge 5\\,mJy, we can deduce important properties about the fainter sub-mm population. Averaging our sub-mm undetected sample (inverse variance weighted) reveals a mean flux of S$_{850}=0.8\\pm0.3$\\,mJy, suggesting that many of these $\\sim$50\\% of our $I>25$ and S$_{1.4}>40$\\,$\\mu$Jy radio sample are still fairly luminous sub-mm sources. They likely form a continuous distribution with the S$_{850}$\\cge 5\\,mJy sample, lie at similar or lower redshifts (Table~1), and comprise roughly 10\\% of the blank field SCUBA source counts from 1--5\\,mJy. This leaves a large portion of sub-mm sources fainter than 5\\,mJy that are not subsumed in our present $I>25$ radio sample. The first possibility is that these fainter sub-mm sources are largely contained in our present radio surveys, but are actually optically bright ($I<25$). Bright LBGs are known to emit at the 1--2\\,mJy level (Chapman et al.~2000a, Peacock et al.~2000), so they must contribute some fraction of the missing sources. This is consistent with the high SFR deduced by Steidel et al.~1999, who apply a large dust correction to their results. Our Arp\\,220 SED model suggests that these sources would be detected in our present radio survey almost out to $z=3$. However, the HDF-SCUBA results (Hughes et al.~1998, Peacock et al.~2000) also show directly that a significant fraction of the 2\\,mJy sources are not associated with bright LBGs. Although deeper radio observations with optically faint counterparts may quickly recover this population, it also remains a possibility that these sources represent high redshift ($z>4$) protogalaxies with L$_{\\rm bol} < 10^{12}$\\Lsun, which would remain undetected in the radio to significantly deeper flux limits. So by pushing to fainter radio limits, it is rather unclear what types of objects might be pre-selected. An additional concern with the sub-mm source population is that massive AGN may be heating the dust rather than star formation. We can be reasonably certain that our radio selected sub-mm sources are not driven primarily by AGN for two reasons. Firstly, the sources are spatially resolved with a median of about 2\\arcsec\\ in the radio using the Merlin interferometer at a resolution of 0.2\\arcsec\\ (Richards 2000, Muxlow et al.~2000), corresponding to $\\sim$1\\,kpc at $z=2$ for our adopted cosmology. If the radio and associated sub-mm emission were emanating from such a compact active nucleus (AGN), it would appear unresolved even at this fine resolution. Secondly, the sub-mm sources have recently been shown to have little or no overlap with X--ray sources as observed with the Chandra satellite (e.g. Fabian et al.~2000, Hornschemeier et al.~2000, Barger et al.~2001). As scenarios which would obscure even the X--ray emission are improbable, the implication is that most bright sub-mm sources are in fact driven by star formation. Assuming then that our radio-selected sources are driven primarily by star formation, it is appropriate to use our data to estimate a contribution to the comoving star formation rate density (SFRD). In Fig.~2, we integrate over our measured counts in three redshift bins, and divide by the effective volume of our detected sources, to represent the sub-mm SFRD as a function of redshift. Since we have assumed the far-IR/radio relation in the redshift estimates, our SFRD estimates can be derived from either wavelength in a manner similar to BCR. We plot our new points as solid squares (sub-mm detected objects) and a solid hexagon (sub-mm undetected objects). The very high redshift open square ($z>3$) represents the Hubble Flanking Fields sub-mm sources undetected in the radio from BCR and Borys et al.~2001. We then compare optically selected sources (open triangles) at redshifts $11$, using the Steidel et al.~(1999) prescription (factor $\\sim$5 for $z>2$). This is roughly in accord with the expected correction from the Chapman et al.~(2000b) sub-mm measurement of 33 LBGs with large expected star formation rates. Lower redshift radio selected sources with bright optical counterparts, as analysed by Haarsma et al.~(2000) are also plotted (red open circles) requiring no correction factor for dust obscuration. Sub-mm sources fainter than 5\\,mJy likely begin to merge with optically selected samples (e.g. Adelberger \\& Steidel 2000). However, our $I>25$ radio selected population is truly an orthogonal population to those discovered in optical surveys, even for S$_{850}$\\cle 5\\,mJy. We can then confidently sum the optically selected (but uncorrected for dust) ($I<25$) and sub-mm ($I>25$) points from $z=$1--4 to arrive at a conservative lower limit to the total SFRD of all presently known objects (plotted as large stars with arrows). This lower limit can be compared with the dust corrected optical points, which is still a rather uncertain procedure. We have therefore recovered the majority of the bright ($>5$\\,mJy) sub-mm sources with a statistically significant sample, selected based on the microJy radio emission with extremely faint optical counterparts. We can state with some assurance that the bulk of bright sub-mm selected sources represent a highly dust obscured star forming population which would be very difficult to identify in optically based surveys. Our redshift analysis has also demonstrated that this is not because the sub-mm sources are at very high redshifts ($z>>3$). These sources may represent the epoch in which the most massive spheroid galaxies were being formed through merging fragments in cluster environments. Indeed, the recent identification of prodigious sub-mm emission with highly over-dense cluster cores at $z\\sim3$ and $z\\sim3.8$ (Chapman et al.~2000c and Ivison et al.~2000 respectively) suggests that the most luminous members of our sub-mm source sample may highlight similar such regions. Pushing our study to fainter sub-mm and radio flux limits will facilitate our understanding of the transition and overlap between these ultra-luminous high$-z$ star formers (which may evolve into the most massive spheroids in the present epoch) and the less massive galaxies selected in the optical through the Lyman-break technique (Steidel et al.~1999)." }, "0011/astro-ph0011316_arXiv.txt": { "abstract": "\\noindent We present analytic solutions of Maxwell equations in the internal and external background spacetime of a slowly rotating magnetized neutron star. The star is considered isolated and in vacuum, with a dipolar magnetic field not aligned with the axis of rotation. With respect to a flat spacetime solution, general relativity introduces corrections related both to the monopolar and the dipolar parts of the gravitational field. In particular, we show that in the case of infinite electrical conductivity general relativistic corrections due to the dragging of reference frames are present, but only in the expression for the electric field. In the case of finite electrical conductivity, however, corrections due both to the spacetime curvature and to the dragging of reference frames are shown to be present in the induction equation. These corrections could be relevant for the evolution of the magnetic fields of pulsars and magnetars. The solutions found, while obtained through some simplifying assumption, reflect a rather general physical configuration and could therefore be used in a variety of astrophysical situations. ", "introduction": "The investigation of the influence of strongly curved spacetimes on the properties of electromagnetic fields has an interest of its own which is further increased when these effects could be related to a rich observable phenomenology. This coupling between general relativistic effects and electromagnetic fields is expected to be particularly important in the vicinity of neutron stars which are among the most relativistic astrophysical objects and are characterized by very intense magnetic fields (Lamb 1991, Glendenning 1996). A number of different observations indicate that in young neutron stars the surface magnetic field strengths are of the order of \\hbox{$10^{11}-10^{13}$ G.} In some exceptional cases, as those of magnetars, magnetic field strengths \\hbox{$\\ge 5 \\times 10^{14}$ G} are considered responsible for the phenomenology observed in soft gamma-ray repeaters (Duncan \\& Thompson 1992, Thompson \\& Duncan 1995). Older neutron stars, observed as recycled pulsars and low mass X-ray binaries, show instead surface magnetic fields that are much weaker $\\le 10^{10}$ G suggesting that these are subject to a decay, even if it is still difficult to establish whether the decay is due to accretion (Geppert \\& Urpin, 1994; Konar \\& Bhattacharya, 1997) or to other processes. In the case of isolated neutron stars, the possibility of magnetic field decay as a result of accretion does not arise, but there are still a number of different ways in which the energy stored in the magnetic can be lost. This can happen either through the emission of electromagnetic (dipole) radiation, through Ohmic decay, through ambipolar diffusion, or through more complicated effects such as ``Hall cascades'' (see Goldreich and Reisenneger 1992 for a review). The investigation of these scenarios requires combined efforts. On one hand, there is the search for a more precise description of the microphysics of the processes involved, some of which are still not well quantified. On the other hand, attention is paid to a more realistic description of the gravitational effects on the properties of the electromagnetic fields in highly curved spacetimes and this is also the motivation of this work. The investigation of the general relativistic corrections to the solution of Maxwell equations in the spacetime of a relativistic star has a long history. The initial works of Ginzburg \\& Ozernoy (1964), Anderson \\& Cohen (1970) and of Petterson (1974) on the stationary electromagnetic fields in a Schwarzschild spacetime have revealed that the spacetime curvature produces magnetic fields which are generally stronger than their Newtonian counterparts (see also Wasserman \\& Shapiro 1983 for a subsequent derivation). Sengupta (1995) has reconsidered this problem and also looked for a general relativistic expression for the electric field in the Schwarzschild background of a neutron star. As we will discuss in Section \\ref{srst_es} the method used in his derivation is not entirely correct and the results obtained for the electric field are not solutions of Maxwell equations. More recently, Sengupta has also considered the problem of the Ohmic decay rate in a Schwarzschild spacetime (Sengupta, 1997). His approach is strictly valid only for the region of spacetime external to the star as it does not provide a correct general relativistic description of the electromagnetic fields internal to the star. Within these approximations, however, Sengupta (1997) has pointed out that the effects of intense gravitational field seem to decrease the overall decay rate by a couple of orders of magnitude. The same problem has also been considered in more detail by Geppert, Page and Zannias (2000). Their analysis was aimed at a mathematically consistent solution of Maxwell equations also in the spacetime region internal to the star and makes therefore use of a generic metric for a non-rotating relativistic star. Their results, while confirming a decrease in the typical decay time for the magnetic field, also show that the decay time is smaller but comparable with the one found in flat spacetime. The general relativistic effects induced by the rotation of the star were first investigated by Muslimov \\& Tsygan (1992) in the slow rotation approximation. A similar approach was also used by Muslimov and Harding (1997) for the electromagnetic fields external to a rotating magnetized star. Their analysis refers to a charge filled magnetosphere and represents the relativistic extension of the Goldreich-Julian model. Using a different derivation, Prasanna and Gupta (1997) have also investigated the properties of the electromagnetic fields in the magnetosphere of a relativistic rotating neutron star, with special attention being paid to the dynamics of charged test particles. We here extend and unify all of the above investigations by considering the solution of Maxwell equations in the internal and external background spacetime of a slowly rotating magnetized relativistic star. The star is considered isolated and in vacuum, with a dipolar magnetic field which is not assumed aligned with the axis of rotation. The purpose of this paper is threefold. Firstly, we want to extend previous results to the most general case of a misaligned rotator, providing for this case also the form of the electric field. Secondly, we want to discuss the possible role played by frame dragging effects in the Ohmic decay for an isolated neutron star and estimate its importance. Thirdly, we wish to clarify a few important aspects of the solution of Maxwell equations in the gravitational field of a relativistic star that, when overlooked, have led to incorrect solutions (Sengupta 1995, Prasanna and Gupta 1997). Finally, by providing a rather general solution to the problem (although truncated at the lowest order in the expansion of the angular dependence) we offer a compact reference from which all of the previous results can be easily found in the appropriate limits and which could have practical astrophysical applications. The paper is organized as follows: in Section \\ref{meq} we write the general relativistic Maxwell equations in the metric of a slowly rotating star and the form they assume when the electromagnetic fields are those measured in the orthonormal frame of zero angular momentum observers. In Section \\ref{ss} we find the stationary solutions (i.e. solutions in which the infinite conductivity of the medium prevents a variation in time of the star's magnetic moment) to Maxwell equations outside and inside the misaligned rotating star. For this we consider first the problem in Newtonian gravity and we then extend the results to general relativity within the slow rotation approximation. Section \\ref{nss} is devoted to the equivalent problem, but in the case in which the magnetic field is not supposed stationary. There, we derive the basic induction equations for the evolution of the inner stellar magnetic field of a misaligned rotating star. Section \\ref{conclusion} contains our conclusions and the prospects of future developments. A number of appendices provide further details about some of the calculations carried out in the main part of the paper. In particular, Appendix A summarizes the components of the electromagnetic tensor in a coordinate basis and in a locally orthonormal tetrad, while Appendix B shows the derivation of the radial eigenfunctions for the electromagnetic fields in terms of Legendre's equation. Appendix C shows the explicit expressions for the surface charges and currents and, finally, Appendix D contains an alternative and equivalent derivation of the equations for the time evolution of magnetic field in terms a vector potential. Throughout, we use a space-like signature $(-,+,+,+)$ and a system of units in which $G = 1 = c$ (However, for those expressions with an astrophysical application we have written the speed of light explicitely.). Greek indices are taken to run from 0 to 3 and Latin indices from 1 to 3; covariant derivatives are denoted with a semi-colon and partial derivatives with a comma. ", "conclusions": "\\label{conclusion} We have presented analytic general relativistic expressions for the electromagnetic fields internal and external to a slowly-rotating magnetized neutron star. The star is considered isolated and in vacuum, but no special assumption is made on the orientation of the dipolar magnetic field with respect to the rotation axis. The solutions to Maxwell equations have been considered both for an infinite and for a finite electrical conductivity. In the first case, corresponding to stationary magnetic fields, we have shown that the general relativistic corrections due to the dragging of reference frames are not present in the form of the magnetic fields but emerge only in the form of the electric fields. In particular, we have shown that the frame-dragging provides an additional induced electric field which is analogous to the one introduced by the rotation of the star in the flat spacetime limit. In the case of finite electrical conductivity, on the other hand, corresponding to decaying magnetic fields, we have shown that corrections due both to the spacetime curvature and to the dragging of reference frames can be found in the induction equation. An interesting result obtained in this regime is that the rotation of the star eliminates the degeneracy in the components of the induction equation which remain therefore distinct. Furthermore, rotation and dipole misalignment do not eliminate the angular dependence in the induction equation and, as a result, an initially dipolar magnetic field might evolve towards a different configuration during its decay. Because of their complexity, the evolution equations found for the magnetic field require a numerical integration which will discuss in detail in a forthcoming work (Rezzolla et al. 2000). There, we will also present direct comparisons between the flat and the curved spacetime solutions and quantify more precisely the importance of the general relativistic corrections. One of the relevant aspects of the solutions presented in this paper is that they provide a lowest order analytic form for the electromagnetic field in the spacetime of a slowly rotating misaligned dipole subject to assumptions which, while giving simplifications, allow the major features of a realistic solution to be seen. In this sense, they reflect a rather general physical configuration and could therefore be used in a variety of astrophysical situations." }, "0011/astro-ph0011498_arXiv.txt": { "abstract": "{ The EPIC focal plane imaging spectrometers on XMM-Newton use CCDs to record the images and spectra of celestial X-ray sources focused by the three X-ray mirrors. There is one camera at the focus of each mirror; two of the cameras contain seven MOS CCDs, while the third uses twelve PN CCDs, defining a circular field of view of 30$^{\\prime}$ diameter in each case. The CCDs were specially developed for EPIC, and combine high quality imaging with spectral resolution close to the Fano limit. A filter wheel carrying three kinds of X-ray transparent light blocking filter, a fully closed, and a fully open position, is fitted to each EPIC instrument. The CCDs are cooled passively and are under full closed loop thermal control. A radio-active source is fitted for internal calibration. Data are processed on-board to save telemetry by removing cosmic ray tracks, and generating X-ray event files; a variety of different instrument modes are available to increase the dynamic range of the instrument and to enable fast timing. The instruments were calibrated using laboratory X-ray beams, and synchrotron generated monochromatic X-ray beams before launch; in-orbit calibration makes use of a variety of celestial X-ray targets. The current calibration is better than 10\\% over the entire energy range of 0.2 to 10 keV. All three instruments survived launch and are performing nominally in orbit. In particular full field-of-view coverage is available, all electronic modes work, and the energy resolution is close to pre-launch values. Radiation damage is well within pre-launch predictions and does not yet impact on the energy resolution. The scientific results from EPIC amply fulfil pre-launch expectations. ", "introduction": "The EPIC instrument on XMM-Newton \\citep{jansen01} provides focal plane imaging and spectrometry for the three X-ray telescopes. Each telescope has an objective comprising a nested, 58 shell, Wolter 1 X-ray mirror \\citep{aschenbach01}, of focal length 7.5 metres, and geometric effective area 1500 cm$^{2}$; there is one EPIC at the focus of each telescope. Two of the telescopes are fitted with the X-ray gratings of the Reflection Grating Spectrometer \\citep{denherder01}. The gratings divert 50\\% of the flux out of the EPIC beams; with allowance for structural obscuration, 44\\% of the original flux reaches two of the EPIC cameras; these contain MOS CCDs \\citep{short98} and are referred to as the MOS cameras. The third telescope has an unobstructed beam; the EPIC instrument at the focus of this telescope uses PN CCDs \\citep{struder01} and is referred to as the PN camera. All three cameras have an identical forward section that contains a filter wheel, door, calibration source, radiation shielding, the interface to the spacecraft focal plane bulkhead, and the internal bulkhead that forms part of the camera vacuum enclosure. The rear part of each camera that contains the CCDs and the cooling system is different in construction for the MOS (figure~\\ref{mos}) and PN cameras. EPIC also includes the EPIC Radiation Monitor System, to record the ambient proton and electron flux \\citep{boer96}. It provides warning of a radiation flux increase to provide for automatic shut down of the instrument. This paper describes the common items and the MOS cameras, while an accompanying paper, by \\cite{struder01} describes the PN camera. \\begin{figure} \\centering \\includegraphics[width=7cm]{Complete_MOS.ps} \\caption{The long conical radiators of the MOS enable the radiating surface to reach the plane of the spacecraft thermal shield to avoid parasitic heat-loads. } \\label{mos} \\end{figure} ", "conclusions": "" }, "0011/astro-ph0011384_arXiv.txt": { "abstract": "The redshift evolution of the galaxy cluster temperature function is a powerful probe of cosmology. However, its determination requires the measurement of redshifts for all clusters in a catalogue, which is likely to prove challenging for large catalogues expected from XMM--Newton, which may contain of order 2\\,000 clusters with measurable temperatures distributed around the sky. In this paper we study the {\\em apparent} cluster temperature, which can be obtained without cluster redshifts. We show that the apparent temperature function itself is of limited use in constraining cosmology, and so concentrate our focus on studying how apparent temperatures can be combined with other X-ray information to constrain the redshift. We also briefly study the circumstances in which non-thermal spectral features can give redshift information. ", "introduction": "Considerable attention has been devoted to the study of the evolution of the galaxy cluster temperature function with redshift, which promises to be an extremely powerful probe of the density parameter (Frenk et al.~1990; Oukbir \\& Blanchard 1992; Viana \\& Liddle 1996; Eke, Cole \\& Frenk 1996). Even very small numbers of high-mass, high-redshift clusters can rule out the critical-density paradigm; indeed, several authors claim that they have already done so (Henry 1997; Bahcall \\& Fan 1998; Eke et al.~1998) though this remains controversial (Sadat, Blanchard \\& Oukbir 1998; Reichart et al.~1999; Viana \\& Liddle 1999). In order to fully apply this method, the cluster masses must be accurately determined and the usual technique is to use the gas temperature as measured from the X-ray emission. In addition the cluster redshift is required, both to place it correctly in the evolutionary sequence, and because the redshift is needed to convert the apparent temperature into the actual cluster temperature. For existing catalogues of clusters for which the temperatures could be estimated, obtaining the redshifts proved a manageable task, as the number of clusters with sufficient photon counts to allow temperature determination was small. This is set to change with observations by the XMM--Newton (hereafter just XMM) satellite; in a recent paper (Romer et al.~1999) we showed that a planned serendipitous cluster survey which will analyze all XMM--EPIC frames suitable for serendipitous cluster detection, {\\sc Xcs}\\footnote{See {\\tt www.xcs-home.org}\\ \\ for further details.}, may contain as many as 10\\,000 galaxy clusters of temperature 2 keV and above, of which around 2\\,000 may have sufficient photon counts to allow the temperatures to be accurately estimated without further X-ray observations. Given that these will be distributed more or less randomly across the sky, follow-up to obtain spectroscopic redshifts represents a substantial task. The main focus of this paper is on the use of the full available X-ray information to optimize the follow-up efficiency onto the high-redshift population. Although a survey like the {\\sc Xcs} is likely to contain around two thousand clusters with high enough photon counts to permit an accurate temperature estimate, the thermal bremsstrahlung spectrum only gives the apparent temperature \\begin{equation} T_{{\\rm app}} = \\frac{T}{1+z} \\,, \\end{equation} with the true temperature not being known until the redshift is determined. For very luminous clusters this degeneracy may be broken by visible spectral lines such as the Iron K line complex at $7$ keV, but this will be challenging for most clusters and we defer discussion of this possibility until the end of the paper. In this paper we discuss several aspects of apparent temperatures and the estimation of cluster redshifts from X-ray data. Apparent temperatures of clusters were first discussed by Oukbir \\& Blanchard (1997) in the context of the ROSAT all-sky survey. They also noted the curious point that even in the absence of redshifts the apparent temperature should still be a good estimator of relative cluster masses; for example in a critical-density model scaling laws predict $M \\propto T^{3/2}/(1+z)^{3/2} \\propto T_{{\\rm app}}^{3/2}$. We focus on issues of apparent temperatures relevant to the XMM satellite. First of all, we analyze whether the apparent temperature function (that is, the number density of clusters observed above a given apparent temperature $T_{{\\rm app}}$) might in itself prove a useful probe of cosmology. The answer will be that it proves of limited use, demonstrating the importance of determining cluster redshifts at the earliest possible stage. In that light, we go on to consider how other X-ray observables can be combined with the apparent temperatures in order to constrain the redshifts, particularly with a view to eliminating low-redshift clusters from the follow-up candidate list. The main observables are the angular size and the apparent luminosity of the clusters, and it is the latter which proves powerful in combination with the apparent temperature. We also briefly study {\\sc xspec} spectral simulations to assess the likelihood of redshift determination from X-ray spectral lines. ", "conclusions": "In this paper we have studied various aspects of apparent cluster temperatures relevant to XMM observations. The apparent temperature function will be readily derived from XMM data, but will primarily be useful in constraining the degenerate combination familiar from low-redshift cluster number density studies, $\\sigma_8 \\Omega_0^{-\\alpha}$ with $\\alpha \\sim 0.5$, and is unlikely to usefully probe the density parameter itself. To fully exploit a galaxy cluster catalogue, cluster redshifts are essential, and we have studied how other X-ray properties can be combined with the apparent temperature to select follow-up candidates efficiently. Assuming point source contamination can be recognized from the high-resolution imaging, it appears that the cluster flux, when combined with the apparent temperature, will yield a good indication of the cluster redshift, especially once early XMM observations have been used to calibrate the relation. Finally, we remark that further useful information towards estimating the redshifts may come from imaging data in the optical and infra-red once cluster candidates have been identified in the X-ray. A good example is the K band luminosity of the brightest cluster galaxy, which exhibits a tight correlation with redshift for X-ray luminous clusters, as shown by Collins \\& Mann (1998) and Burke, Collins \\& Mann (2000). In many cases it may also be possible to carry out photometric redshifting of cluster galaxies, for example using Sloan Digital Sky Survey and VISTA data." }, "0011/hep-ex0011001_arXiv.txt": { "abstract": "We present the first direct experimental evidence for the charge excess in high energy particle showers predicted nearly 40 years ago by Askaryan. We directed bremsstrahlung photons from picosecond pulses of 28.5 GeV electrons at the SLAC Final Focus Test Beam facility into a 3.5 ton silica sand target, producing electromagnetic showers several meters long. A series of antennas spanning 0.3 to 6 GHz were used to detect strong, sub-nanosecond radio frequency pulses produced whenever a shower was present. The measured electric field strengths are consistent with a completely coherent radiation process. The pulses show 100\\% linear polarization, consistent with the expectations of Cherenkov radiation. The field strength versus depth closely follows the expected particle number density profile of the cascade, consistent with emission from excess charge distributed along the shower. These measurements therefore provide strong support for experiments designed to detect high energy cosmic rays and neutrinos via coherent radio emission from their cascades. ", "introduction": "During the development of a high-energy electromagnetic cascade in normal matter, photon and electron scattering processes pull electrons from the surrounding material into the shower. In addition, positrons in the shower annihilate in flight. The combination of these processes should lead to a net 20-30\\% negative charge excess for the comoving compact body of particles that carry most of the shower energy. G. A. Askaryan\\cite{Ask62} first described this effect, and noted that it should lead to strong coherent radio and microwave Cherenkov emission for showers that propagate within a dielectric. The range of wavelengths over which coherence obtains depends on the form factor of the shower bunch---wavelengths shorter than the bunch length suffer from destructive interference and coherence is lost. However, in the fully coherent regime the radiated energy scales quadratically with the net charge of the particle bunch, and at ultra high energies the resulting coherent radio emission may carry off a significant fraction of the total energy in the cascade. The plausibility of Askaryan's arguments combined with more recent modeling and analysis\\cite{Mark86,ZHS92,Alv96,Alv97,Alv98} has led to a number of experimental searches for high energy neutrinos by exploiting the effect at energies from $\\sim 10^{16}$ eV in Antarctic ice\\cite{fri96,Bes99} up to $10^{20}$ eV or more in the lunar regolith, using large ground-based radio telescopes\\cite{Zhe88,Dag89,Han96,Gor99}. Radio frequency pulses have been observed for many years from extensive air showers\\cite{Jel66,Feg68}. However, it has been shown\\cite{KL66,All71} that the most likely source of this emission is a form of Lorentz-boosted dipole radiation from geomagnetic charge separation in the air shower, rather than the Askaryan effect. Thus neither the charge asymmetry nor the resulting coherent Cherenkov radiation has ever been observed. In a previous paper\\cite{P1} we have described initial efforts to measure the coherent radio-frequency (RF) emission from electron bunches interacting in a solid dielectric target consisting of 360~kg of silica sand. That study, done with relatively low-energy electrons (15 MeV), demonstrated the presence of coherent radiation in the form of extremely short and intense microwave pulses detectable over a wide frequency range. These results, while useful for understanding the coherent RF emission processes from relativistic charged particles, could not directly test the development of a shower charge excess, as Askaryan predicted. Also, because the particles were charged, passage of the beam through any interface induced strong RF transition radiation (TR), which obscured the presence of Cherenkov radiation (CR). We report here on measurements made at the Stanford Linear Accelerator Center (SLAC) in which the use of high-energy photons, rather than low-energy electrons, has enabled us to clearly observe microwave Cherenkov radiation from the Askaryan effect. The electromagnetic showers thus produced in our target yielded strong, coherent, sub-ns RF pulses, consistent in every way observed with the predictions. As we will show here, our results conclusively support the reported sensitivity of the experiments noted above. ", "conclusions": "We have demonstrated that the radiation we have observed is coherent, and 100\\% linearly polarized. The plane of polarization coincides with the plane containing the antenna and shower axis. The radiation is pulsed with time durations much shorter than the inverse bandwidth of our antennas. The strength of the pulse is strongly correlated to the size (in particle number) of the shower region that appears to produce it. All of these observed characteristics are consistent with the hypothesis that what we have observed is coherent Cherenkov radiation. Because of the strong correlation with the shower profile, and the physical constraint that a shower with no net excess charge cannot radiate, we conclude that excess charge production along the shower is the source of the propagating charge in the silica dielectric which leads to the CR observed. These conclusions are strengthened by the fact that the pulse strength is consistent with Monte Carlo predictions for such showers. Our observations are inconsistent with radiation from geomagnetic charge separation as observed in extensive air showers. The most striking evidence for this is the fact that the plane of polarization is clearly aligned with the shower axis rather than the local geomagnetic dip angle ($62^{\\circ}$). Given the approximate east-west orientation of the shower axis, boosted dipole radiation from geomagnetic charge separation should produce an electric field polarization with significant components orthogonal to what we have observed. We note that in all cases our measurements are likely to be made in near-field conditions, and we have not attempted to correct for these effects. Recent studies\\cite{Bun00,Alv00} have begun to treat these issues, but are not straightforward to apply in our case. In any case, near field effects should generally {\\em decrease} the measured field strengths relative to far field measurements. The total energy of our cascades is as high as $10^{19}$ eV, but these showers consist of the superposition of many lower-energy showers. Higher-energy effects\\cite{Alv98,Alv99} that elongate the shower during its development are not present. These effects all tend to increase the total tracklength of the shower, at the expense of a lower instantaneous particle number density. The net effect is that the total radiated power is expected to be approximately conserved, but the angular spectrum can change significantly, with a predicted sharpening of the Cherenkov angular distribution for high energy showers. We have not measured the angular spectrum of the radio pulses, and thus any scaling of our results to high energy showers should consider corrections for these effects, which are likely to increase the peak field strength in far-field measurements. Since the field strength scales linearly in shower energy and inversely with distance from the source, extrapolations to determine the energy threshold of existing experiments is straightforward, after correcting for the differences in material properties. For Antarctic ice experiments, the use of the existing simulations \\cite{ZHS92,Alv98,fri96,Alv00} appears completely justified. For experiments observing the lunar regolith\\cite{Zhe88,Dag89,Han96,Gor99}, silica sand shares many similarities with the lunar surface material, and the expected cascade energy threshold, by direct scaling of our results, is $\\sim 2 \\times 10^{19}$ eV, and may be somewhat lower depending on the angular intensity effects discussed above. We conclude that, in combination with our previous measurements of coherent RF transition radiation~\\cite{P1}, these results have established a firm experimental basis for radio-frequency detection of high energy cascades in solid media, either through interaction within a dielectric (for CR), or via passage through dielectric interfaces (for TR). Above cascade energies of $\\sim 10^{16}$~eV, these secondary emission processes become dominant over others (for example, optical Cherenkov or fluorescence emission) in the number of quanta produced\\cite{Mark86}. Thus experiments designed to exploit this effect in the detection of ultra-high energy particles can now be pursued with even greater confidence." }, "0011/astro-ph0011047_arXiv.txt": { "abstract": "We have obtained HST-NICMOS observations of five of M31's most metal rich globular clusters: G1, G170, G174, G177 \\& G280. For the two clusters farthest from the nucleus we statistically subtract the field population and estimate metallicities using $K$-$(J-K)$ color-magnitude diagrams (CMDs). Based on the slopes of their infrared giant branches we estimate [Fe/H] $=-1.22\\pm0.43$ for G1 and $-0.15\\pm0.37$ for G280. We combine our infrared observations of G1 with two epochs of optical HST-WFPC2 $V$-band data and identify at least one LPV based on color and variability. The location of G1's giant branch in the $K$-$(V-K)$ CMD is very similar to that of M107, indicating a higher metallicity than our purely infrared CMD: [Fe/H]$\\sim -0.9\\pm0.2$. For the three central clusters, which are too compact for accurate cluster star measurements, we present integrated cluster magnitudes and field CMDs. The $K$-band luminosity functions (LFs) of the upper few magnitudes of G1 and G280, as well as for the fields surrounding all clusters, are indistinguishable from the LF measured in the bulge of our Galaxy. This indicates that these clusters are very similar to Galactic clusters, and at least in the surrounding fields observed, there are no significant populations of young luminous stars. For the field surrounding G280, we estimate the metallicity to be $-1.3$ from the slope of the giant branch, with a spread of $\\sigma_{[Fe/H]} \\sim 0.5$ from the width of the giant branch. Based on the numbers and luminosities of the brightest giants, we conclude that only a small fraction of the stars in this field could be as young as 2 Gyr, while the majority have ages closer to 10 Gyr. ", "introduction": "\\label{sec:introduction} Globular clusters (GC) occupy a very special position in modern astrophysics. They provide the most stringent tests of stellar evolution theory, represent ideal templates for stellar population synthesis studies, allow age dating of galaxies with unrivaled precision, etc. The GC families of the Milky Way and its satellite galaxies, the Magellanic Clouds and the Fornax dSph galaxy, have been thoroughly studied from both the ground and space, especially with HST. The next nearest major GC family belongs to M31. To study individual stars in GC systems much more distant than M31 will require the next generation of ground or space based telescopes. The GC systems of the the Milky Way and M31 appear to be quite similar in luminosity and color ranges \\citep{BBBF1987, EW1988, HBK1991, FPC1980b, DGMC1997, SBA1996}. However, some of M31's bulge globulars appear to have line strengths as strong as those of giant ellipticals, suggesting metallicities considerably greater than any known Galactic globular \\citep{BJSA1992, JAB1992, Jab1997}. \\citet{BFGK1984} has also noted that M31's GCs appear to follow different H$\\beta$ and CN correlations with Mg$_2$ with respect to Galactic globulars; they attributed this to the whole M31 GC family being systematically younger. Renzini (1986) pointed out that other interpretations were also possible, such as the chemical enrichment history of the two spheroids having proceeded with slightly different time scales, resulting in different element ratios. HST observations of M31's GCs in the optical (with the goal of stellar photometry) started soon after the first refurbishing mission, with both FOC and WFPC2 \\citep{FBCC1996, RMFN1996, AGLB1996}, and now include some of the most metal rich clusters \\citep{JCMS2000}. The study of metal rich M31 globulars complements similar studies of Galactic bulge globulars within this still poorly known, yet crucial part of age-metallicity space. Near-IR observations are essential to study the brightest giants in an old, metal rich population, especially to determine bolometric luminosities. Metal rich globular and bulge giants, which are the brightest bolometrically and in the near-IR, are in contrast, many magnitudes fainter at optical wavelengths due to severe molecular blanketing. In extreme cases, $(V-K)$ can be as great as $\\sim 10$ at the top of the AGB \\citep{FW1987, GOMR1998} and these stars are likely to have escaped detection in M31 even with WFPC2. Since the main sequences of the old populations in M31 are currently out of reach, one is forced to appeal to an alternative age indicator. Theory predicts that the highest luminosity reached on the AGB is a function of age \\citep{IR1983}, a prediction extensively verified by observations of clusters in the Magellanic Clouds \\citep{MA1986, FMB1990}. In metal poor Galactic globulars, no stars are brighter than the theoretical RGB tip, as expected for a $\\sim 14$ Gyr population. However, the AGB of more metal rich clusters ([Fe/H]$>-1$) extends $\\sim 1$ magnitude above the RGB \\citep{FPC1983, FE1988, GRO1997}, as it does in the Galactic bulge \\citep{FW1987}. For clusters, though, these bright stars are all LPVs. For both clusters and bulge population these luminous stars are now generally ascribed to high metallicity, rather than to young age \\citep{FW1987, GRO1997}. Hence, the presence of stars brighter than the RGB tip does not guarantee an intermediate age; consideration of their color, luminosity {\\it and} frequency in the parent population is required before drawing conclusions about ages \\citep{Ren1993}. In Cycle 7 we proposed to obtain NICMOS $JHK$ images of 5 metal rich globular clusters in M31. From these observations, we planned to achieve the following scientific goals: (1) explore the upper end of the GC luminosity function; (2) derive independent estimates for the metallicity of the M31 clusters and their adjacent fields from the slope and location of the NIR RGB \\citep{KFTP1995}; (3) determine the frequency of luminous RGB and AGB stars (including LPVs) per unit luminosity; (4) compare the cluster results with observations of M31 field stars (5) compare the properties of the luminous stars in M31's metal rich clusters to those in the Galactic bulge and Galactic globulars; (6) integrate these results with optical photometry and spectroscopy and explore implications for stellar evolution theory and the interpretation of the integrated light of distant galaxies. One critical issue deserves special attention if any of these goals are to be attained: the effect of stellar crowding. We \\citep[][hereafter Paper I]{SFFJ2001}, have carefully analyzed the effects of blending on our NICMOS data. Through the creation of completely artificial clusters, we have calculated threshold- and critical-blending limits for each cluster and surrounding field. These limits determine the proximity to each cluster where reliable photometry can be obtained. These simulations allow us to quantify and correct for the effects of blending on the GB slope and width at different surface brightness levels. This paper is organized as follows. Section \\ref{sec:observations} presents the reasons for selecting each cluster, and the details of the observations. Section \\ref{sec:data_reduction} describes the reduction procedures, and gives a brief summary of the procedures and results on blending from Paper I. Section \\ref{sec:photometry} presents the integrated photometry of the clusters, the CMDs and luminosity functions of the clusters and their surrounding fields, and metallicity estimates for G1, G280, and the G280 field. Our conclusions are summarized in Section \\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} We first present surface brightness profiles of all clusters, and with the blending analysis presented in Paper I, determine radial photometric limits for each cluster. We then give integrated photometry of all clusters except G1, which was not fully on the detector. For the G1 cluster, we present the infrared CMD, and estimate the metallicity as [Fe/H]$= -1.22\\pm0.43$ from the slope of the giant branch. Based on the width of the giant branch, which shows no significant spread in color over what is expected from measurement errors alone $(\\sigma_{(J-H)}=0.06)$, we conclude that there is no significant metallicity spread in the cluster. We combine our infrared observations of G1 with two epochs of optical $V$-band HST-WFPC2 data, revealing that several of the brightest stars in the cluster are LPVs. The shape and position of the GB in the $K$-$(V-K)$ CMD are similar to that of M107, indicating a metallicity of [Fe/H]$=-0.9\\pm0.2$. However since the infrared GB slope technique uses such a small range in luminosity, we place more weight on the higher value from the optical-infrared CMD. For the G280 observations, we divide the frame into cluster and field at $5''$ from the cluster center. We statistically subtract the field population from the cluster, and present both the cluster and field CMDs. Fitting the giant branch, we find a cluster metallicity of [Fe/H]$= -0.15\\pm0.37$. As in G1, we see no evidence for a metallicity spread in the cluster based on the width of the GB $(\\sigma_{(J-H)}=0.07)$. Fitting the GB of the G280 field, we find a metallicity of $-1.3 \\pm 0.6$. The large error on the metallicity is indicative of the large color spread, which we estimate to be $\\sigma_{[Fe/H]} \\sim 0.5$ dex from the width of the GB. This is not surprising, since this field has contributions from both the disk (85\\%) and bulge (15\\%). What is surprising is that, in this field which is 85\\% disk, we see no obviously bright, young stars. Using the brightest star in the field as the tip of the AGB, at $M_{bol}=-5$, we estimate an age of $\\sim 5$ Gyr. However, if the disk component were this young, we would expect to see $\\sim 30$ stars brighter than $M_{bol}=-4$, but we see only 7. A more likely scenario is that there are just a few young disk stars in the field, while the majority of the disk population is closer to $\\sim 10-15$ Gyr, thus lowering the AGB tip to $M_{bol} \\sim -4.5$. The three central clusters, G170, G174 \\& G177, are all too compact to extract cluster star photometry. We thus present the field CMDs and luminosity functions without trying to separate the cluster and field star contributions. The surface brightness of the G170 field is within acceptable limits, so we perform a linear fit to the GB, but do not try to estimate the metallicity, as the blending correction would be uncomfortably large. The G174 and G177 fields, on the other hand, are both above the threshold-blending surface brightness limit. This implies that measurements of all but the brightest stars in these fields are potentially affected by blending, so we refrain from even fitting their GBs. Finally we presented the cluster (G1 and G280) and field luminosity functions with the LF measured in Baade's Window. The luminosity functions of G1 and G280 both have a sharp bright-end cutoff at $M_K \\sim -6.5$, consistent with observations of Galactic globulars. The fields surrounding the clusters have LFs which are indistinguishable from that measured in the Galactic bulge. Thus, at least in the fields observed, there is no significant population of young luminous stars in the bulge of M31. \\" }, "0011/astro-ph0011271_arXiv.txt": { "abstract": "{ We report on the first deep X-ray survey with the {\\em XMM-Newton} observatory during the performance verification phase. The field of the Lockman Hole, one of the best studied sky areas over a very wide range of wavelengths, has been observed. A total of $\\sim$ 100 ksec good exposure time has been accumulated. Combining the images of the {\\em European Photon Imaging Camera} (EPIC) detectors we reach a flux limit of 0.31, 1.4 and $2.4 \\times 10^{-15}~{\\rm erg}~{\\rm cm}^{-2}~{\\rm s}^{-1}$, respectively in the 0.5-2, 2-10, and 5-10 keV band. Within an off-axis angle of 10 arcmin we detect 148, 112 and 61 sources, respectively. The log(N)-log(S) relation in the three bands is compared with previous results. In particular in the 5-10 keV band these observations present the deepest X-ray survey ever, about a factor 20 more sensitive than the previous {\\em BeppoSAX} observations. Using X-ray spectral diagnostics and the set of previously known, spectroscopically identified {\\em ROSAT} sources in the field, the new sources can be classified. {\\em XMM-Newton} detects a significant number ($\\sim$ 40\\%) of X-ray sources with hard, probably intrinsically absorbed X-ray spectra, confirming a prediction of the population synthesis models for the X-ray background. ", "introduction": "\\label{sec:intr} Deep X-ray surveys indicate that the cosmic X-ray background (XRB) is largely due to accretion onto supermassive black holes, integrated over cosmic time. In the soft (0.5-2 keV) band 80-90\\% of the XRB flux has been resolved using {\\em ROSAT} and recent {\\em Chandra} surveys (Hasinger et al. \\cite{hasi98a}, Mushotzky et al. \\cite{mushotzky00}, Giacconi et al. \\cite{giacconi00}). In the harder (2-10 keV) band 25-30\\% of the background have been resolved in {\\em ASCA} and {\\em BeppoSAX} surveys (Ueda et al. \\cite{ueda98}, Cagnoni et al. \\cite{cagnoni98}, Giommi et al. \\cite{giommi00}), and more than 60\\%, when the recent {\\em Chandra} surveys are included. Surveys in the very hard (5-10 keV) band have been pioneered using {\\em BeppoSAX} and resolve about 30\\% of the XRB (Fiore et al. \\cite{fiore99}). Those X-ray surveys with a high degree of completeness in optical spectroscopy find predominantly Active Galactic Nuclei (AGN) as counterparts of the faint X-ray source population (Bower et al. \\cite{bower96}, Schmidt et al. \\cite{schmidt98}, Zamorani et al. \\cite{zamorani99}, Akiyama et al. \\cite{akiyama00}), mainly X-ray and optically unobscured AGN (type-1 Seyferts and QSOs) but also a smaller fraction of obscured AGN (type-2 Seyferts). Spectroscopic identifications of the {\\em BeppoSAX} and {\\em Chandra} surveys are still far from complete, however a mixture of obscured and unobscured AGN seems to be the dominant population in these samples, too (Fiore et al. \\cite{fiore00}, Barger et al. \\cite{barger00}, Giacconi et al. \\cite{giacconi00}). The most recent AGN X-ray luminosity function, derived from the {\\em ROSAT} surveys, shows evidence for luminosity-dependent density evolution and indicates a constant QSO space density at redshifts $2 < z < 4$ (Hasinger \\cite{hasi98b}, Miyaji et al. \\cite{miyaji00}) unlike optical QSO luminosity functions (Schmidt et al. \\cite{schmidt95}, Fan et al. \\cite{fan00}). The X-ray observations are consistent with population synthesis models based on unified AGN schemes (Setti \\& Woltjer \\cite{setti89}, Madau et al. \\cite{madau94}, Comastri et al. \\cite{comastri95}, Gilli et al. \\cite{gilli99}), which explain the hard spectrum of the X-ray background by a mixture of absorbed and unabsorbed AGN, folded with the corresponding luminosity function and its cosmological evolution. According to these models most AGN spectra are heavily absorbed and about 80\\% of the light produced by accretion will be absorbed by gas and dust (Fabian et al. \\cite{fabian98}). However, these models are not unique and contain a number of hidden parameters, so that their predictive power remains limited (e.g. Hasinger \\cite{hasi00}). In particular they require a substantial contribution of high-luminosity obscured X-ray sources (type-2 QSOs), which so far have not been detected in sufficient quantities (see the discussion in Halpern et al. \\cite{halpern99}). The large throughput and the unprecedented hard X-ray sensitivity of the telescopes aboard the recently launched {\\em XMM-Newton} observatory (hereafter {\\em XMM}; Jansen et al. \\cite{jansen01}) will ultimately yield spectra of the faint X-ray sources and constrain the evolution of their physical properties, in particular the X-ray absorption. Here we present results of the first deep survey taken with {\\em XMM} in the {\\em Lockman Hole}, one of the best studied sky areas at all wavelengths. This paper concentrates on the X-ray data analysis. We show combined images from the EPIC pn-CCD (Str\\\"uder et al. \\cite{strueder01}) and MOS CCD cameras (Turner et al. \\cite{turner01}) and the derived source counts in different energy bands. With the help of X-ray colour-colour diagrams and the previously identified sources in this field we show that it is possible to obtain a coarse source classification based on {\\em XMM} data alone. ", "conclusions": "\\label{sec:disc} We have shown the first log(N)-log(S) relations based on the {\\em XMM-Newton} observatory. Given the still existing systematic uncertainties, the data is fully consistent with the {\\em ROSAT} and {\\em Chandra} source counts in the 0.5-2 keV band. This demonstrates on one hand that the combination of EPIC detectors is not yet confusion limited in a 100 ksec observation, on the other hand that cosmic variance between different fields does not affect the source counts significantly at the currently achieved flux levels, at least not in the soft band. In the 2-10 keV band there is an inconsistency of about 40\\% between the two recent {\\em Chandra} datasets by Mushotzky et al. \\cite{mushotzky00} and Giacconi et al. \\cite{giacconi00}, the latter one having a lower normalisation. The new {\\em XMM} data are consistent with the Giacconi et al. log(N)-log(S), maybe even somewhat flatter, and clearly confirm a break in the slope compared to the quasi-Euclidean behaviour at brighter fluxes. In the 5-10 keV band the {\\em XMM} data go more than an order of magnitude deeper than the previous {\\em BeppoSAX} counts (Fiore et al. \\cite{fiore99}). There is so far relatively little deviation from a Euclidean slope and the data is fully consistent with the predictions from recent population synthesis models for the X-ray background (Gilli et al. \\cite{gilli99}). Adding up the source counts, we resolve about 60\\% of the 5-10 keV X-ray background. The diagnostic power of {\\em XMM} lies in its wide energy band and its unprecedented sensitivity in the hard band. With the help of X-ray colour-colour diagrams and the ``training set'' of about 60 previously identified {\\em ROSAT} sources in the same field it is possible to characterise the new XMM sources as typically harder, probably intrinsically absorbed sources. A small number of objects with similar X-ray colours has already been identified in the deepest ROSAT survey (Lehmann et al. \\cite{lehmann00}), they are type-2 Seyferts or unidentified objects with extremely red optical/NIR colours ($R-K > 5$). The new XMM source population is therefore very likely dominated by obscured AGN, as predicted by the AGN population synthesis models for the X-ray background." }, "0011/astro-ph0011101_arXiv.txt": { "abstract": "The dynamic orchestration of starbirth activity in the starburst-ringed galaxy M94 (NGC 4736) is investigated using images from the Ultraviolet Imaging Telescope (FUV-band), Hubble Space Telescope (NUV-band), Kitt Peak 0.9-m telescope (H$\\alpha$, R, and I bands), and Palomar 5-m telescope (B-band), along with spectra from the International Ultraviolet Explorer and Lick 1-m telescopes. The wide-field UIT image shows FUV emission from (a) an elongated nucleus, (b) a diffuse inner disk, where H$\\alpha$ is observed in {\\it absorption}, (c) a bright inner ring of H II regions at the perimeter of the inner disk (R = 48$''$ = 1.1 kpc), and (d) two 500-pc size knots of hot stars exterior to the ring on diametrically opposite sides of the nucleus (R = 130$''$ = 2.9 kpc). The HST/FOC image resolves the NUV emission from the nuclear region into a bright core and a faint 20$''$-long ``mini-bar'' at a position angle of 30 deg. Optical and IUE spectroscopy of the nucleus and diffuse inner disk indicates a $\\sim$10$^{7-8}$ yr-old stellar population from low-level starbirth activity blended with some LINER activity. Analysis of the H$\\alpha$, FUV, NUV, B, R, and I-band emission along with other observed tracers of stars and gas in M94 indicates that most of the star formation is being orchestrated via ring-bar dynamics involving the nuclear mini-bar, inner ring, oval disk, and outer ring. The inner starburst ring and bi-symmetric knots at intermediate radius, in particular, argue for bar-mediated resonances as the primary drivers of evolution in M94 at the present epoch. Similar processes may be governing the evolution of the ``core-dominated'' galaxies that have been observed at high redshift. The gravitationally-lensed ``Pretzel Galaxy'' (0024+1654) at a redshift of $\\sim$1.5 provides an important precedent in this regard. ", "introduction": "Star-forming rings or ``pseudorings'' are common to early and intermediate-type spiral galaxies (cf.\\markcite{Athan85} Athanassoula \\& Bosma 1985; \\markcite{buta95b}Buta, Purcell, \\& Crocker 1995; \\markcite{bc96}Buta \\& Combes 1996), including our own Milky Way galaxy (cf. \\markcite{gm82}Gusten \\& Mezger 1982; \\markcite{Clemens88}Clemens, Sanders, \\& Scoville 1988; \\markcite{Waller90a}Waller 1990a). Such ring-like accumulations of gas and associated starbirth activity may have helped to build the inner parts of many primeval disk galaxies (\\markcite{Friedli95}Friedli \\& Benz 1995), as exemplified by the recent discovery of a starburst-ringed galaxy at $z \\sim 1.5$ (\\markcite{Colley96}Colley, Tyson, \\& Turner 1996; \\markcite{Tyson97}Tyson et al. 1997). The formation and maintenance of these starburst rings are often attributed to orbital resonances with rotating bar or ``oval'' asymmetries in the stellar disks (cf. \\markcite{Combes94}Combes 1994, \\markcite{Byrd94}Byrd et al. 1994; \\markcite{Combes95}Combes et al. 1995; \\markcite{bc96}Buta \\& Combes 1996 and references therein). However, other dynamical mechanisms --- including gravitational instabilities (\\markcite{Elmegreen92}Elmegreen 1992, \\markcite{Elmegreen94}1994; \\markcite{Kenney97}Kenney \\& Jogee 1997), outward propagating star formation (\\markcite{Walker88}Walker, Lebofsky, \\& Rieke 1988; \\markcite{Waller92}Waller, Gurwell, \\& Tamura 1992), and even radially-driven pileups from nuclear outbursts (Waller et al. 1992; \\markcite{Tenorio97}Tenorio-Tagle et al. 1997) --- may play significant roles in orchestrating some of the starburst rings that are observed. \\footnote{Collisionally-induced ``ring galaxies'' such as the Cartwheel Galaxy are thought to be morphologically and dynamically distinct from the more common ``ringed galaxies'' considered herein (cf. \\markcite{Athan85}Athanassoula \\& Bosma 1985; \\markcite{Marcum92} Marcum et al. 1992).} As the closest early-type spiral galaxy of low inclination, M94 (NGC 4736) has received concentrated attention from both observers and theorists. This (R)SA(r)ab-type galaxy (\\markcite{rc3}de Vaucouleurs et al. 1991) is noted for its inner ring of ongoing starburst activity (R $\\approx$ 45$''$), oval stellar distribution at intermediate radius (R $\\approx$ 220$''$) (cf. \\markcite{Mulder93}Mulder \\& van Driel 1993; \\markcite{Mulder95}Mulder 1995; \\markcite{Mollenhoff95} Mollenhoff, Matthias, \\& Gerhard 1995), and outer stellar ring near its de Vaucouleurs radius (R$_{25}$ $\\approx$ 330$''$). {\\bf Figure 1} (extracted from the Digital Sky Survey using the SkyView utility [McGlynn, Scollick, \\& White 1996])\\footnote{NASA's SkyView facility (http://skyview.gsfc.nasa.gov) was developed and is maintained under NASA ADP Grant NAS5-32068 at NASA's Goddard Space Flight Center.} shows the outermost portions of M94, highlighting the oval disk and outer pseudoring. The inner starburst ring is a prominent source of H$\\alpha$, H I, and CO emission (\\markcite{smith91}Smith et al. 1991; \\markcite{mulder93}Mulder \\& van Driel 1993; \\markcite{Gerin91} Gerin, Casoli, \\& Combes 1991). The discovery of compact thermal \\& nonthermal radio sources in the ring (\\markcite{Duric88}Duric \\& Dittmar 1988) indicates the presence of dense H II regions and young SNRs. The ring's velocity field can be described by a combination of circular rotation with velocities of order 200 km/s and residual non-circular motions of order 15 km/s (\\markcite{Mulder95}Mulder 1995) to 25 km/s (\\markcite{buta88}Buta 1988), depending on the adopted inclination and major axis position angle. Interior to the ring, the bright bulge and inner disk show twisted isophotes at red and near-IR wavelengths, indicative of a weak bar-like distortion (\\markcite{Beckman91}Beckman et al. 1991; \\markcite{Shaw93}Shaw et al. 1993; \\markcite{Mollenhoff95} Mollenhoff et al. 1995). FIR and CO observations interior to the ring reveal a rich ISM with gas surface densities exceeding that of the ring (\\markcite{Smith94}Smith \\& Harvey 1994; \\markcite{Garman86}Garman \\& Young 1986; \\markcite{Gerin91}Gerin et al. 1991; \\markcite{Wong00}Wong \\& Blitz 2000). Optical spectroscopy of the nuclear region yields LINER-type emission lines along with absorption lines from the circumnuclear stellar population, consistent with an early main-sequence stellar turnoff (A4--A7) and corresponding age of $\\sim$500 Myr (\\markcite{Pritchett77} Pritchett 1977; \\markcite{Keel83}Keel 1983; \\markcite{Taniguchi96}Taniguchi et al. 1996). Further support for a young central population comes from NIR spectroscopy which shows deep CO absorption bands from red giant and asymptotic giant branch stars of similar age (\\markcite{Walker88}Walker et al. 1988). These authors have proposed an outward propagating mode of star formation, whereby NGC 253, M82, M94, and M31 represent increasingly evolved versions of the same starbursting sequence. Although the kinematics of the ring show very little evidence for outward {\\it expanding motions} (contrary to prior claims of bulk expansion [\\markcite{kruit74}van der Kruit 1974; \\markcite{kruit76}1976]), they also do not preclude a scenario for {\\it radially propagating} star formation. Other investigators have modeled the inner and outer rings in terms of {\\it resonant dynamics} mediated by bar or ``oval'' potentials interior to the rings (\\markcite{Gerin91}Gerin et al. 1991; \\markcite{Shaw93} Shaw et al. 1993; \\markcite{Mollenhoff95}Mollenhoff et al. 1995; \\markcite{Mulder96}Mulder \\& Combes 1996), with the observed non-circular motions resulting from dispersion orbits near the Lindblad resonances (\\markcite{buta88}Buta 1988). In this paper, we present and discuss new observational clues to the dynamical mechanisms governing the star formation in M94. The ultraviolet images obtained by the Ultraviolet Imaging Telescope, in particular, reveal hitherto unrecognized patterns of recent star formation whose presence lends further support to the hypothesis of galaxy evolution via bar-mediated resonances. The various imaging and spectroscopic observations and reductions are described in Section 2. The resulting FUV, NUV, H$\\alpha$, R, and I-band emission morphologies are presented and compared in Section 3. Radial intensity profiles and other photometric comparisons are discussed in Section 4. UV and optical spectroscopy of the inner disk and nucleus is presented in Section 5. Kinematic properties and inferred dynamical scenarios are considered in Section 6. Our summary of the wavelength-dependent morphological, spectro-photometric, and dynamical properties of M94 appears in Section 7, wherein evolutionary implications are discussed. In the following Sections, we assume a distance to M94 of 4.6 (75/H$_{\\circ}$) Mpc, based on the galaxy's recession velocity of 345 km/s with respect to the Local Group (\\markcite{Sandage81} Sandage \\& Tammann 1981). The corresponding spatial scale is 22.3 pc/arcsec. Unresolved sources imaged by the HST/FOC provide additional constraints on the distance, as discussed in Section 4. We adopt a nominal inclination of 40$^{\\circ}$ and major-axis position angle of 120$^{\\circ}$, while recognizing that both of these quantities may vary significantly with radius and with measuring technique (e.g. morphological vs. kinematic determinations) (\\markcite{Bosma77}Bosma, van der Hulst, \\& Sullivan 1977; \\markcite{buta88}Buta 1988; \\markcite{Mulder93}Mulder \\& van Driel 1993; \\markcite{Mulder95}Mulder 1995; \\markcite{Mollenhoff95}Mollenhoff et al. 1995; \\markcite{Wong00}Wong \\& Blitz 2000). ", "conclusions": "Through UV-Optical imaging and spectroscopy, we have found new evidence for bar-mediated resonances as the primary drivers of evolution in M94 at the present epoch. Our observational results include evidence for \\noindent\\hang(1.) A 450-pc long nuclear ``mini-bar'' at both optical and near-UV wavelengths. \\noindent\\hang(2.) An inner disk with diffuse FUV emission in concentric arcs that do not match the fine-scale structures or reddened structures at visible wavelengths. Since $H\\alpha$ is observed in absorption here, the UIT/FUV image represents the first view of this non-ionizing but relatively young disk component. \\noindent\\hang(3.) UV-Optical colors and spectral indices in the nucleus and inner disk that indicate B and A-type stars in the presence of modest extinction (A$_V$ $\\le$1 mag) along with some LINER activity from the nucleus itself. \\noindent\\hang(4.) A 2.2 kpc diameter starbursting ring at the perimeter of the inner disk that is bright at FUV, H$\\alpha$, and radio-continuum wavelengths. The level of starbirth activity in this inner ring rivals the levels observed in starbursting irregular galaxies such as NGC 1569 and NGC 4449. The inferred star formation rate within the ring and inner disk amounts to 1.5 M$_{\\odot}$ yr$^{-1}$ --- sufficient to build up the stellar mass of the inner disk and bulge in $\\sim$10$^{10}$ yr. \\noindent\\hang(5.) No detectable radial offsets between the H$\\alpha$ and FUV rings, thus indicating a 35 km/s speed limit to any outward or inward propagating star formation in the ring, if such a mode is present. \\noindent\\hang (6.) Two 500-pc size FUV-emitting knots exterior to the ring on diametrically opposite sides of the nucleus. The bisymmetric knots and starburst ring appear to be especially prominent parts of a complex spiral arm structure, as revealed in a spatially-filtered B-band image. \\noindent\\hang (7.) The starburst ring, bi-symmetric knots, oval disk, and outer pseudo-ring as signposts of resonant dynamics in the disk of M94. More specifically, the radii of these features match those of various orbital resonances, given a pattern speed of 35 km s$^{-1}$ kpc$^{-1}$ at our adopted distance and inclination. These orbital resonances are most likely driven by some combination of the nuclear mini-bar and oval distortion in the disk. \\noindent\\hang (8.) A shallow minimum of gravitational stability at the radius of the starburst ring that extends inward into the inner disk. Although too broad to explain the discrete starburst ring, the shallow minimum may help to explain the 10$^7-10^8$-yr old stellar population interior to the ring. \\vskip 12pt Although we can set a limit on the speed of outward or inward propagating star formation in the ring, we cannot preclude the existence of such a mode. At a propagation speed of 35 km/s, a wave initiated in the nucleus could traverse the inner disk to the radius of the current starburst ring in only 31 Myr. Therefore, it is possible that the $10^7-10^8$-yr old stellar population detected in the inner disk is the result of such an outward propagating wave. The striking difference in emission morphologies at FUV and red wavelengths provides further support for the starburst ring being a transient phenomenon which does not persist at any one radius for very long. Either these resonant phenomena come and go, as the oval distortions undergo secular evolution, or their operating radii migrate in response to other dynamical influences on the stars and gas (cf. \\markcite{Combes94}Combes 1994; \\markcite{Combes95}Combes et al. 1995; \\markcite{Friedli95}Friedli \\& Benz 1995). Otherwise, one must invoke strong radial inflows of stars from the starburst ring to populate the inner disk and bulge, a feat requiring unusual circumstances -- e.g. mergers. The results reported herein may have important implications with regard to observations of the most distant observable galaxies. At redshifts of 1--5, the 2-kpc diameter starburst ring in M94 would subtend angles of only (0.7$''$ -- 1.0$''$)H$_{\\circ}$/75 in an Einstein-de Sitter Universe (q$_{\\circ}$ = 1/2) and (0.3$''$ -- 0.2$''$)H$_{\\circ}$/75 in an open (Milne) universe (q$_{\\circ}$ = 0) (cf. \\markcite{Narlikar83}Narlikar 1983). The UV-bright nuclear rings evident in NGC 1097, NGC 1317, NGC 1433, NGC 1512, NGC 2997, NGC 4321, and NGC 5248 (\\markcite{maoz95}Maoz et al. 1995; \\markcite{maoz96}Maoz et al. 1996; \\markcite{kuchinsky00}Kuchinsky et al. 2000; \\markcite{marcum00}Marcum et al. 2000) would subtend even smaller angles at the same redshifts. Moreover, nuclear rings tend to have higher FUV surface brightnesses than their larger counterparts -- the inner ring in M94 being a remarkable exception. Therefore, some of the ``core-halo'' morphologies that are evident at high-redshift in the restframe FUV (cf. \\markcite{Giavalisco97}Giavalisco et al. 1997) may, in fact be marginally-resolved representations of galaxies with starburst rings in their centers. Gravitationally-lensed galaxies are fortuitously magnified, enabling resolutions of their structure at high S/N. An important precedent in this regard is the gravitationally-lensed ``Pretzel Galaxy'' which lies behind the galaxy cluster 0024+1654 at an estimated redshift of 1.2 -- 1.8 (\\markcite{Colley96}Colley et al. 1996; \\markcite{Tyson97}Tyson et al. 1997). Detailed reconstructions of the multiply-lensed galaxy show a clear annular morphology on a scale of several kpc. If M94 and other nearby ringed galaxies can be used as current-epoch analogues, the ``Pretzel Galaxy'' and perhaps other marginally-resolved ``core-halo'' galaxies at high redshift may represent youthful inner disks and bulges growing under the organizing influence of oval or bar asymmetries(\\markcite{Friedli95}Friedli \\& Benz 1995; \\markcite{Waller97}Waller et al. 1997). Conversely, if evidence for starburst rings at high redshift proves to be sparse, then massive inner disks featuring ring-bar dynamics have yet to form in most systems, or starbursting bulges are masking their presence." }, "0011/astro-ph0011323_arXiv.txt": { "abstract": "{ The \\object{Coma} cluster of galaxies was observed with XMM-Newton in 12 partially overlapping pointings. We present here the resulting X-ray map in different energy bands and discuss the large scale structure of this cluster. Many point sources were found throughout the observed area, at least 11 of them are coincident with bright galaxies. We also give a hardness ratio map at the so far highest angular resolution obtained for a cluster of galaxies. In this map we found soft regions at the position of bright galaxies, little variation in the central 15 arcmin, but some harder regions north of the line \\object{NGC 4874} -- \\object{NGC 4889}. ", "introduction": "The \\object{Coma} cluster of galaxies was one of the performance verification targets of XMM-Newton. The aim was to verify XMM-Newton's ability to map large extended X-ray sources. This is not an easy task because errors can occur in several areas. The telescope vignetting plays a major role when overlapping the same region of an extended source, measured at different off-axis angles. Because of the vignetting, the surface brightness of a region measured off axis seems to be lower than the same region measured on axis. The reason for that is the increasing loss of mirror collecting area with increasing off-axis angle, which in addition is energy dependent. By de-vignetting the different observations and normalizing the observations to the obser\\-ving times the surface brightnesses of the same source regions are corrected and must show the same number of counts/arcmin$^2$/sec. If done correctly, detector-edges or regions of detector gaps should not be visible as inhomogenities in the surface brightness of the merged observations. \\begin{table*}[t] \\caption[ ]{Journal of Observations} \\begin{tabular}{llccccc} \\hline \\rule{0mm}{4mm} & Name of & & & \\multicolumn{3}{c}{pn-Camera observing times (ksec)}\\\\ \\multicolumn{1}{c}{Date} & Observation & RA(2000) & DEC(2000) & planned & \\multicolumn{1}{c}{performed} & \\multicolumn{1}{c}{effective} \\\\ \\hline \\rule{0mm}{4mm} 2000 May 29 \\ & Coma center & 12 59 46.7 & 27 57 00 & 15.0 & 22.7 & 15.4 \\\\ 2000 June 21/22 & Coma 1 & 12 56 47.7 & 27 24 07 & 25.0 & 29.2 & 24.0 \\\\ 2000 June 11/12 & Coma 2 & 12 57 42.5 & 27 43 38 & 25.0 & 34.8 & 34.7 \\\\ 2000 June 27/28 & Coma 3 & 12 58 32.2 & 27 24 12 & 25.0 & 27.0 & 14.3 \\\\ 2000 June 23 & Coma 4 & 13 00 04.6 & 27 31 24 & 25.0 & 20.0 & (11.4) \\\\ 2000 May 29 & Coma 5 & 12 59 27.5 & 27 46 53 & 20.0 & 25.0 & 10.0 \\\\ 2000 June 12 & Coma 6 & 12 58 50.0 & 27 58 52 & 20.0 & 14.8 & 3.7 \\\\ 2000 June 24 & Coma 7 & 12 57 27.7 & 28 08 41 & 25.0 & 8.1 & (3.4) \\\\ not observed yet & Coma 8 & 13 01 25.6 & 27 43 53 & 26.0 & -- & -- \\\\ 2000 June 12 & Coma 9 & 13 00 32.7 & 27 56 59 & 20.0 & 26.3 & 21.3 \\\\ 2000 June 22 & Coma 10 & 12 59 38.4 & 28 07 40 & 20.0 & 26.3 & 16.6 \\\\ 2000 June 24 & Coma 11 & 12 58 36.5 & 28 23 56 & 25.0 & 32.1 & 15.9 \\\\ not observed yet & Coma 12 & 13 01 50.2 & 28 09 28 & 25.0 & -- & -- \\\\ not observed yet & Coma 13 & 13 00 36.5 & 28 25 15 & 25.0 & -- & -- \\\\ 2000 June 22 & Coma bkgd & 13 01 37.0 & 27 19 52 & 30.0 & 22.7 & 20.1 \\\\ \\hline \\end{tabular} \\end{table*} The backgrounds in the image are problematic. The part of the background that is produced by particles inside the detector is not distinguishable from X-rays, is not vignetted, and is in general distributed homogeneously over the detector face (except for the lower end of the spectral range and around the detector intrinsic emission lines around 8 keV, see Briel et al. \\cite{ub00}). Hence this component must be subtracted before de-vignetting is applied. The extragalactic X-ray background is vignetted in the same manner as diffuse emission from an extended source and may therefore be de-vignetted in the same manner. But there is a third component in the background, the low energy protons, that are again indistinguishable from X-rays. They come through the mirror system and show a different vignetting compared to X-rays (see Str\\\"uder et al. \\cite{ls01}). In addition, their spectrum and their flux is time variable. Because of the different vignetting, it is essential to subtract a ``proton image\" from the measured image before de-vignetting is applied, especially when the surface brightness of the proton induced background is in the same order of magnitude as the surface brightness of the extended source. \\object{Coma} was chosen because it is a well studied source at all wavelength ranges. With a diameter of the main cluster, which is several times larger than the FOV of the EPIC cameras of about 26 arcmin, it is necessary to mosaic the cluster with XMM-Newton if the large scale structure and dynamics of the cluster are to be investigated. Although \\object{Coma} was long believed to be the archetype of a relaxed cluster of galaxies, the X-ray image taken during the Rosat All Sky Survey revealed the complex large scale structure, especially the subgroup around \\object{NGC 4839} probably falling into the main cluster (Briel et al. \\cite{ub92}). Subsequent pointed observations with Rosat showed that \\object{Coma} has irregular structure on different scales and was most probably formed by hierarchical clustering of several subunits, which are still visible in their X-ray emission (White et al. \\cite{sw93}). Observations with ASCA, and also the pointed Rosat observation, showed evidence of temperature structure in \\object{Coma} (Honda et al. \\cite{hh96}; Briel \\& Henry \\cite{ub98}; Donnelly et al. \\cite{rd99}; Watanabe et al. \\cite{mw99}). The interpretation of the temperature structure favors the ongoing merging of the subgroup around \\object{NGC 4839} with the main cluster, a scenario which is still debated. In this letter we will present the result of merging 12 partially overlapping observations of the \\object{Coma} cluster with the EPIC pn-CCD-detector on board XMM-Newton. In Section 2 we will describe the observations and present the data reduction and the method of merging the individual observations. In Section 3 we will discuss some of the many point sources in the image, the large scale structure of the X-ray surface brightness distribution, and the hardness ratio map of the inner cluster region. ", "conclusions": "The \\object{Coma} cluster is the nearest very rich cluster of galaxies. It is probably the best studied cluster at all wavelengths and is one of the brightest extragalactic X-ray sources in the sky. Until now, the deepest X-ray observation of \\object{Coma} is that of White et al. (\\cite{sw93}) consisting of 4 exposures of about 19 ksec each by the ROSAT PSPC. The EPIC pn observations reported here are of comparable length, thus are $\\sim$8 times deeper, while simultaneously extending over $\\sim$3 times wider energy range and with $\\sim$4 times better angular resolution. Perhaps not surprisingly given the better spatial re\\-solution and deeper observation, the most noticable difference between our XMM images in Figures 1 and 2 compared to Figure 2 of White et al. (\\cite{sw93}) is the myriad of newly detected point-like sources. Dow \\& White (\\cite{kd95}) discuss the $\\sim$25 point sources in the ROSAT observations of \\object{Coma}. There are at least 75 point sources in the smaller solid angle we have analyzed. We have made a first attempt to identify these sources by overlaying our X-ray map on an optical image of the region and then consulting NED when there appears to be a coincidence. Many bright galaxies seem to be X-ray sources including: \\object{NGC 4827}, \\object{4839}, \\object{4858}, \\object{4860}, \\object{4874}, \\object{4889}, \\object{4911}, and \\object{4921}, and \\object{IC 4040} and \\object{4051}. However some faint galaxies, with magnitudes from 18 to 20, could also be identifications based on positional coincidence. At least three quasars, \\object{QSO 1256+281}, \\object{QSO 1258+280}, and \\object{QSO 1259+281}, are detected. However, the bulk of the sources are unidentified in our initial analysis. The \\object{NGC 4911} and \\object{NGC 4839} groups contain concentrations of point sources at our sensitivity. Dow \\& White (\\cite{kd95}) and Neumann et al. (\\cite{dn01}) discuss the particularly interesting morphology of the \\object{NGC 4839} group. The large-scale diffuse emission in the soft band shown in Figure 1 is very similar to the ROSAT observations of White et al. (\\cite{sw93}) in the same band. The X-ray emission of the \\object{Coma} cluster is lumpy near its center but becomes smoother at larger radii. An additional lump, apparent in the ROSAT observation but not discussed by White et al. (\\cite{sw93}), is directly NE of \\object{NGC 4839}. It may be yet another group on its way into the main cluster to the ENE. This lump may be preceeding the \\object{NGC 4839} group from the direction of \\object{A1367} since calculations of cluster growth indicate that the groups that merge to form clusters flow along filaments connecting the clusters. The large-scale diffuse hard emission is very similar to the soft, as Figure 2 shows. Because the calibration of the spectral response of the extended full frame mode of the pn-CCD is not yet available, we do not yet want to perform spatially resolved spectral fits. We can, however, make a hardness ratio map using wide bands, which we show in Figure 3. While these maps sacrifice spectral details, such as temperatures and abundances, they are the highest angular ($\\sim$40 arcsec) and spatial ($\\sim$26 kpc for H$_{\\mbox{o}}$ = 50 km s$^{-1}$ Mpc$^{-1}$) resolution spectral maps ever obtained for a cluster of galaxies. Using the conversion from hardness ratio to temperature, given in Fig. 3, we have checked our resulting mean temperature in some regions which are known from other measurements. For the over-all temperature within a radius of 10 arcmin, centered on \\object{NGC 4874}, we determined a temperature of 8.8 keV. The GINGA value is 8.21 $\\pm$ 0.09 keV (Hughes et al. \\cite{jh93}), the ASCA value is 9. $\\pm$ 0.6 keV (Donnely et al \\cite{rd99}) and the XMM-MOS value is 8.25 $\\pm$ 0.10 keV (Arnaud et al \\cite{ma01}). We agree to better than 10\\% with these previously determined values, which would be expected given our understanding of the uncertainties of the calibration and of the background model. In a further check, we determined the temperature around \\object{NGC 4874} within a radius of 0.5 arcmin and within a ring from 1 to 5 arcmin. Our result is 7 and 9.2 keV, the MOS result (Arnaud et al \\cite{ma01}, Fig. 8) is 6.6 and 8.6 keV respectively, again in agreement within 10\\%. We found the same kind of agreement on other places, taking into account the large area of 3.5 $\\times$ 3.5 arcmin$^2$, used for the MOS data by Arnaud et al. (\\cite{ma01}). The large-scale description of the data in Figure 3 agrees with Arnaud et al. (\\cite{ma01}): there is cool gas to the south-east and hotter gas to the south-west and west. It agrees to some extent with Donnely et al. (\\cite{rd99}), Watanabe et al. (\\cite{mw99}) and Briel \\& Henry (\\cite{ub98}) investigations. In particular the south-east cool region was already identified by Donnely et al. (\\cite{rd99}) and a hot region in the south-west/west direction, but somewhat further away, was found by Watanabe et al. (\\cite{mw99}). There are spectral variations on the smallest scales that we can measure. Softer emission is associated with many compact sources some of which are associated with galaxies as discussed previously. Although the smoothing of the original surface brightness distribution smears out the point sources to some extend, a significant drop in the radial temperature profile within 1 arcmin of \\object{NGC 4874} was also seen by Arnaud et al. (\\cite{ma01}), Fig. 8. The Rosat temperature map of Briel \\& Henry (\\cite{ub98}) also shows cooler regions around the two central galaxies (although other cool regions in the center of their map are not confirmed). If these soft regions, associated with galaxies, are extended, we can postulate three possible explanations. First, they are local potential wells centered on cluster galaxies as is was discussed by Vikhlinin et al. (\\cite{av94}). The shallower potential wells could only trap lower temperature gas. Second, they are the interstellar emission that the cluster galaxy has been able to retain. Third, integrated emission from low mass X-ray binaries which Irwin \\& Sarazin (\\cite{ji98}) have discussed as the source of soft emission from elliptical galaxies. Perhaps the most interesting small scale spectral structures are the isolated hard spots that do not appear to be associated with any obvious point-like X-ray emission. For the most part they are not associated with an optical object in NED. The five hard spots north of the \\object{NGC 4889} -- \\object{NGC 4874} line may be the cause of the hot region in the same general location detected by Donnelly et al. (\\cite{rd99}) in their ASCA observations. It seems that these hot spots are washed out in the temperature map of Arnaud et al. (\\cite{ma01}) because of the larger (3.5$\\times$3.5 armin$^2$) integration boxes. They might have been seen however in the even larger integration region of ASCA, if they are associated with variable sources like QSO's. But we can not entirely eliminate an instrumental cause for some of them because some are aligned parallel to the CCD orientation and on the boarder of the chips." }, "0011/astro-ph0011115_arXiv.txt": { "abstract": "A search for star forming objects belonging to tidal tails has been carried out in a sample of deep H$\\alpha$ images of 16 compact groups of galaxies. A total of 36 objects with H$\\alpha$ luminosity larger than 10$^{38}~erg~s^{-1}$ have been detected in five groups. The fraction of the total H$\\alpha$ luminosity of their respective parent galaxies shown by the tidal objects is always below 5\\% except for the tidal features of HCG~95, whose H$\\alpha$ luminosity amounts to 65\\% of the total luminosity. Out of this 36 objects, 9 star forming tidal dwarf galaxy candidates have been finally identified on the basis of their projected distances to the nuclei of the parent galaxies and their total H$\\alpha$ luminosities. Overall, the observed properties of the candidates resemble those previously reported for the so-called tidal dwarf galaxies. ", "introduction": "Interactions between disk galaxies are known to produce strong morphological disturbances and, if the geometry of the encounter is favorable, to enhance the star formation rates of the disk galaxies. The pioneering work by Tomre \\& Tomre (1972) and a later study by Barnes (1988) supported the original idea by Zwicky (1956) that dwarf galaxies could originate within the tidal tails resulting from a strong interaction between two disk galaxies. Since then, several examples of dwarf galaxies with tidal origin have already been detected in interacting systems in an advanced stage of merging and also around galaxies experiencing less dramatic interactions (Schweizer 1978, Mirabel et al. 1991,1992, Hibbard \\& van Gorkom 1996, Duc \\& Mirabel 1997,1998). The compact group environment has been suggested to be an ideal place for interactions between galaxies given their large spatial densities and their low velocity dispersions. Mendes de Oliveira \\& Hickson (1994) reported morphological signs of interaction in many galaxies of compact groups from Hickson's Catalogue (1982). Other recent studies on the photometric properties of the galaxies in Hickson Compact Groups at different wavelengths resulted contradictory when compared to samples of field galaxies: while CO, FIR, H$\\alpha$ and optical fluxes seem to be similar to those measured for field galaxies (Moles et al. 1994, Verdes-Montenegro et al. 1998, Iglesias-P\\'{a}ramo \\& V\\'{\\i}lchez 1999), radio continuum and 21cm emission are on average lower than for samples of isolated galaxies (Williams \\& Rood 1987, Menon 1995, Huchtmeier 1997). Hunsberger et al. (1998) carried out a search for dwarf galaxy candidates in the close environment of a sample of Hickson Compact Groups, by automated detection of faint objects on deep $R$ band frames. After decontamination for the presence of background galaxies, they found that the luminosity function in compact groups showed an enhancement at the faint luminosity end compared to a normal Schechter function. They concluded that the initial dwarf galaxy population in compact groups is replenished by ``subsequent generations'' formed in the tidal debris of giant galaxy interactions. In this work we present the results of a search for tidal star forming dwarfs in our H$\\alpha$ images of a sample of nearby compact groups (see V\\'{\\i}lchez \\& Iglesias-P\\'{a}ramo 1998, Iglesias-P\\'{a}ramo \\& V\\'{\\i}lchez 1999). Five of the groups were found to show optically prominent tidal tails with star forming objects within them. We emphasize the fact that all the detected objects are H$\\alpha$ emitters, therefore implying that they are physically associated with the galaxies of the group. The main properties of the star forming objects will be presented in \\S\\ref{sample}, as well as a description of their close environment. \\S\\ref{discussion} contains a discussion on their likely evolutionary scenario. ", "conclusions": "} As can be seen from Table~1, the H$\\alpha$ equivalent widths and luminosities of the tidal objects suggest that most of them host very young star formation bursts and are above the luminosity limit for giant H{\\sc ii} regions. The underlying continuum luminosity varies greatly from object to object as shown in Figures~1 to 5. This is the reason for the large range covered by their H$\\alpha$ equivalent width values, rather than the result of different evolutionary stages of the corresponding star formation bursts of the tidal objects. The fraction of the total H$\\alpha$ flux corresponding to tidal objects is typically under 5\\% (see Table~1), except in the remarkable case of the tidal features in HCG~95c, which amount to 65\\% of the total luminosity of the galaxy. We attribute this extraordinarily high fraction of luminosity within the tidal tail to the fact that HCG~95c belongs to a triple system showing a rather complicated interaction sequence (Iglesias-P\\'{a}ramo \\& V\\'{\\i}lchez 1998). The existence of these tidal objects is linked to the possibility that they could subsequently evolve as independent dwarf galaxies, like those observed at the tips of the tidal tails of the Antennae and the Superantennae (Mirabel et al. 1991,1992). Looking to the optical images, only rough morphological similarities between the tails of the Antennae and Superantennae and the tails shown by our galaxies can be found. The Antennae and Superantennae fall in the most dramatic case of interaction of a pair of galaxies, whereas our five groups span a wide range of interaction strengths. At this point, two critical questions arise concerning the possibility that tidal objects can evolve independently as dwarf galaxies: (1) are they distant enough from the parent galaxy so as to be able to escape from the potential well?, and (2) are they massive enough to be self gravitating systems and to be considered dwarf galaxies? Concerning the first question, some authors have claimed (Schweizer 1978; Hibbard \\& van Gorkom 1996) that most of the material ejected during an interaction is accumulated in the tidal tails. N-body simulations suggest that a substantial fraction of this material will slowly fall into the main body of the remnant of the interaction, and only the outermost 20\\% will probably gain enough kinetic energy to be able to evolve independently for a long time (see Hibbard \\& Mihos 1995). Given that no information is available on the velocity fields of the interacting pairs of our sample, and that they show a great variety of interaction patterns, we will adopt a quantitative criterion based on the projected distances of the tidal objects from the nuclei of their parent galaxies in order to select the most likely candidates to escape from the potential well. Ferguson et al. (1998) carried out a deep search for H{\\sc ii} regions well outside the optical limits of a sample of disk galaxies, and found no H{\\sc ii} regions located at a projected galactocentric distance larger than $2 \\times R_{25}$. This result was explained as due to the fact that the gas surface density is below the critical density for star formation at such large radii from the centers of the galaxies. Therefore, we can consider that those star forming regions located that far from the nuclei of their parent galaxies must have been originated by an external mechanism which has removed gas from the inner regions of the parent galaxies. Star forming regions far away from their parent galaxies may become kinematically decoupled from the stellar disks; thus they are very likely escaping the potential well of the parent galaxy and could become good tidal dwarf galaxy candidates. According to Table~\\ref{prop}, {\\em only 9 tidal objects fulfill this criterion} ($R > 2 \\times R_{25}$). Obviously, this is a conservative criterion adopted given the lack of information on the velocity field of the tidal tails. Therefore, we should bear in mind that because of possible projection effects, the number of tidal dwarfs candidates selected is a lower limit since other objects not satisfying this criterion cannot be ruled out to be tidal dwarf galaxies. Concerning the second question posed above, Elmegreen et al. (1993) proposed that clouds as massive as $10^{8}~M_{\\odot}$ can form during interactions of disk galaxies. Under some conditions, these clouds could become dwarf galaxies. In particular, when the mass of the perturber is larger than approximately 1.4 times the mass of the parent galaxy, -- as appears to be the case for the five groups of our sample presenting H$\\alpha$ emitting tidal tails\\footnote{estimated from a comparison of the magnitudes of the parent galaxies with those of the remaining galaxies in the groups} -- these models predict that the outermost clouds are very likely to escape the potential well. Such $10^{8}~M_{\\odot}$ clouds would produce $10^{6}~M_{\\odot}$ starbursts, assuming a value for the efficiency of star formation of 1\\% (see Kennicutt 1998 and references therein). From the observational side, the H$\\alpha$ luminosities reported for the tidal dwarf galaxies of the Antennae (Mirabel et al. 1991) and NGC~5291 (Duc \\& Mirabel 1998), are comparable to the ones derived in the present paper; the H{\\sc i} masses measured for those dwarfs appear to be typically above $10^{8}~M_{\\odot}$. In the same direction, from a large compilation of dwarf galaxies Hunter \\& Gallagher (1986) reported H{\\sc i} masses larger than $10^{8}~M_{\\odot}$ for objects with $L_{\\mbox{\\scriptsize{H}}\\alpha} \\gtrsim 10^{39}~erg~s^{-1}$. Such bright objects are known to be gravitationally bound, according to the well established $L($H$\\alpha$) {\\em versus} $\\sigma$ relationship for H{\\sc ii} galaxies (Terlevich \\& Melnick 1981). The 9 objects selected above show H$\\alpha$ luminosities larger than $10^{39}~erg~s^{-1}$, which on observational grounds seems to be a reasonable lower limit to ensure self gravitation for H{\\sc ii} complexes. The two restrictions imposed above, though conservative, appear reasonable since H{\\sc ii} regions brighter than $10^{39}~erg~s^{-1}$ are found to be located at galactocentric distances lower than about $1.2 \\times R_{25}$ (as derived from the H{\\sc ii} region surveys by Rozas et al. 1996,2000). A naive estimation of the dynamical time needed by our dwarf galaxy candidates to travel a distance equivalent to $R_{25}$ as a consequence of the interaction would be of the order of the fading time for H{\\sc ii} regions. Thus, we are confident that the 9 selected candidates must have been generated during the interaction and do not belong to a previous generation of disk H{\\sc ii} regions. From the arguments presented above, it appears clear that the 9 star forming objects finally selected represent good candidates to satisfy the {\\em escape} and {\\em self gravitation} conditions, and thus they can be included in the group of the so-called tidal dwarf galaxies. Further information on total masses and relative velocity with respect to their parent galaxies is required to confirm that they satisfy both conditions. It seems striking that the ratio between the number of candidate dwarf galaxies to the total number of tidal objects detected, 25\\%, appears consistent with the fraction of the ejected material which is expected to escape according to theoretical models (Hibbard \\& van Gorkom 1995). Since all these candidate tidal galaxies were selected at projected distances larger than $2\\times R_{25}$, we suggest that this last criterion might be a useful indicator of an {\\em effective} escape radius for these systems. Summarizing, as a result of a search performed on a sample of net H$\\alpha$ images of 16 compact groups of galaxies, we have identified 9 star forming tidal objects proposed to be considered dwarf galaxy candidates given their large projected distances from their parent galaxies and their large H$\\alpha$ luminosities. These interesting objects deserve further detailed study, including dynamical and evolutionary aspects which are beyond the scope of the present study." }, "0011/astro-ph0011265_arXiv.txt": { "abstract": " ", "introduction": "The concept of biasing between different classes of extragalactic objects and the background matter distribution was put forward by Kaiser (1984) and Bardeen et al. (1986) in order to explain the higher amplitude of the 2-point correlation function of clusters of galaxies with respect to that of galaxies themselves. In this framework biasing is assumed to be statistical in nature; galaxies and clusters are identified as high peaks of an underlying initially Gaussian random density field. Biasing of galaxies with respect to the dark matter distribution was also found to be an essential ingredient of CDM models of galaxy formation in order to reproduce the observed galaxy distribution (cf. Davies et al. 1985; Benson et al. 2000). The classical approach to study the redshift evolution of bias utilises the ratio of the correlation functions of objects and dark matter, which are assumed to be related via the square of a scale independent bias factor. However, in this study we will use the definition by which the extragalactic mass tracer (galaxies, halos, clusters) fluctuation field, $\\delta_{tr}$, is related to that of the underlying mass, $\\delta_{\\rm m}$, by \\be\\label{eq:1} \\delta_{\\rm tr} = b \\delta_{\\rm m} \\;\\;, \\ee where $b$ is the linear bias factor. Note that the former definition results from the latter but the opposite is not necessarily true. The bias factor may have many dependencies; even assuming that it is scale independent, it necessarily depends on the type of the mass tracer as well as on the epoch $z$, since the fluctuations evolve with time as gravity draws together galaxies and mass. It is evident, therefore, that the bias factor should also depend on the different cosmological models and dark matter content of the Universe (for a recent overview see Klypin 2000). More realistic biasing schemes have been proposed in the literature. Coles (1993) introduced the idea of biased galaxy formation in which galaxies form with a probability given by an arbitrary function of the local mass density. Mann, Peacock \\& Heavens (1998) investigated the properties of different bias models of galaxy distributions that results from local transformations of the present-day density field. The deterministic and linear nature of eq.(\\ref{eq:1}) has been challenged (cf. Dekel \\& Lahav 1999; Tegmark \\& Bromley 1999) and indeed some non-linearity of the biasing relation is necessary to reconcile high biasing with deep voids. Despite the above, the linear biasing assumption is still a useful first order approximation which, due to its simplicity, it is used in most studies of large scale (linear) dynamics (cf. Strauss \\& Willick 1995 and references therein; Branchini et al. 1999; Schmoldt et al. 1999; Plionis et al. 2000). Different studies have indeed shown that the bias factor is a monotonically decreasing function of redshift. An important advancement in the analytical treatment of the bias evolution was the work of Mo \\& White (1996) in which they used the Press-Schechter (1974) formalism and found that in an Einstein-de Sitter universe the linear bias factor evolves strongly with redshift. Using a similar formalism, Matarrese et al. (1997) extended the Mo \\& White results to include the effects of different mass scales (see also Catelan et al 1998). Steidel et al. (1998) confirmed that the Lyman-break galaxies are very strongly biased tracers of mass and they found that $b(z=3.4) \\magcir 6, 4, 2$, for SCDM, $\\Lambda$CDM $(\\Omega=0.3)$ and OCDM $(\\Omega=0.2)$, respectively (see also Giavalisko et al 1998). A similar value for the $\\Lambda$CDM model was obtained by Cen \\& Ostriker (2000) using high resolution Nbody/hydro simulations in which they treated DM, gas as well as star formation. The use of high resolution N-body simulations (cf. Klypin et al. 1996; 1999, Cole et al. 1997 and references therein) have shown that anti-biasing ($b< 1$) should exist at scales $r \\sim 3 - 8 h^{-1}$ Mpc, for the open and flat low-$\\Omega$ models, in contrast with $\\Omega=1$ models, where $b>1$. Colin et al (1999), using high-resolution N-body simulations of SCDM, $\\Lambda$CDM, OCDM and $\\tau$CDM models, which avoid the so called ``overmerging\" problem, found that indeed biasing evolves rapidly with redshift, while Kauffmann et al. (1999) combining semi-analytic models of galaxy formation and N-body simulations has also studied the evolution of clustering in different cosmologies. In this paper we will not indulge in such aspects of the problem but rather, working within the paradigm of linear and scale-independent bias, we will derive the functional form of its redshift evolution in the matter dominated epoch and in all cosmological models. The Einstein de-Sitter case has been studied in the past (cf. Nusser \\& Davis 1994; Fry 1996; Bagla 1998) using the continuity equation, which is a first order differential equation, to derive a solution, $\\propto (1+z)$, valid only for low $z$'s. Our approach is to use the perturbation evolution equation which combines the continuity, the Euler and the Poisson equations and which is a second order differential equation. We should therefore expect to find a further component to the known solution. The paper is organised as follows: in section 2 we discuss the basic models for the linear bias evolution, in section 3 we derive the basic differential equation describing the evolution of the linear bias factor, while in section 4 we present its analytical solution for the different cosmological models and a comparison with previous models and N-body simulation results. Finally, in section 5 we summarise our main results. ", "conclusions": "We have introduced analytical arguments and approximations based on linear perturbation theory and a linear, scale-independent bias between a mass tracer and its underlying matter fluctuation field in order to investigate the cosmological evolution of such a bias. We derive a second order differential equation, the solution of which provides the functional form of the of bias evolution in any Universe. For the case of an Einstein-de Sitter Universe, we find an exact solution which is a linear combination of the known solution $\\propto (1+z)$ (cf. Bagla 1998 and references therein), derived from the continuity equation, and a second term $\\propto (1+z)^{3/2}$ which dominates. This solution once parametrised at two different epochs, compares extremely well with the more sophisticated halo models (cf. Mo \\& White 1996) and with N-body simulations. For the two low-density cosmological models we find exact solutions, albeit only in the high-redshift approximation (where the growing mode of perturbations can be approximated by the Einstein-de Sitter solution). We also derive analytical solutions for two limiting low-density Universes (ie., $\\Omega=0$, $\\Omega_{\\Lambda}=0$ and $\\Omega=0$, $\\Omega_{\\Lambda}=1$)." }, "0011/astro-ph0011053_arXiv.txt": { "abstract": "OO Aql is a rare W UMa-type eclipsing binary in which the two solar-type stars may have only recently evolved into contact. The binary has an unusually high mass ratio (0.84), and a relatively long orbital period (0$\\fd$506) for it spectral type (mid-G). Twelve ultraviolet spectra of OO Aql were obtained in 1988 with the {\\it IUE} satellite, including a series of consecutive observations that cover nearly a complete orbital cycle. Chromospheric activity is studied by means of the Mg~{\\sc ii} h$+$k emission at 2800 {\\AA}. The Mg~{\\sc ii} emission is found to vary, even when the emission is normalized to the adjacent continuum flux. This variation may be correlated with orbital phase in the 1988 observations. It also appears that the normalized Mg~{\\sc ii} emission varies with time, as seen in spectra obtained at two different epochs in 1988 and when compared with two spectra obtained several years earlier. The level of chromospheric activity in OO Aql is less than that of other W UMa-type binaries of similar colors, but this is attributed to its early stage of contact binary evolution. Ultraviolet light curves were composed from measurements of the ultraviolet continuum in the spectra. These were analyzed along with visible light curves of OO Aql to determine the system parameters. The large wavelength range in the light curves enabled a well-constrained fit to a cool spot in the system. ", "introduction": "The eclipsing binary star system OO Aquilae (HD 187183, $V_{\\rm max}=9.2$, $B-V=+0.76$, mid-G) is an unusual W UMa-type binary because it possesses a mass ratio near unity (0.84; Hrivnak 1989). This suggests that the components have only recently (on astronomical timescales) come into contact, since once in contact the stars are expected to evolve toward low mass ratio (Webbink 1976; Vilhu 1982). The orbital period of 0$\\fd$506 is long compared to other G spectral types stars with W UMa-type light curves. This results in an unusually large angular momentum for a cool W UMa-type system, again suggesting that it has only recently evolved into contact (Mochnacki 1981). Thus OO Aql represents a rare, transient phase in the evolution of contact binary star systems. On the basis of his modern radial velocity study using the cross-correlation technique, Hrivnak (1989) carried out a consistent analysis of both the light and velocity curves of OO Aql. He determined the absolute parameters of the two stars with high precision (M$_1$=1.04$\\pm$0.02 M$_{\\sun}$, R$_1$=1.39$\\pm$0.02 R$_{\\sun}$, M$_2$=0.88$\\pm$0.02 M$_{\\sun}$, R$_2$=1.29$\\pm$0.02 R$_{\\sun}$). By comparing its properties with stellar models, Hrivnak presented evidence that the stars have an age of $\\sim$8 Gyr. Thus OO Aql consists of a pair of mid-G stars in contact, which appear to be somewhat more evolved than the Sun. It falls into the subclass of A-type W UMa-type binaries, in which the more massive component is eclipsed at primary minimum; this is unusual among the cooler (G and K spectral type) W UMa-type binaries. W UMa-type binaries of G and K spectral type are known to display a high level of chromospheric activity, as shown in the study of Rucinski (1985). Most contact binaries of G$-$K spectral types have periods of 0$\\fd$25 to 0$\\fd$35 and mass ratios of 0.3 to 0.5. Thus OO Aql is clearly unusual. Another contact binary with a large mass ratio (0.94) is VZ Psc (P=0$\\fd$26, K2$-$5). This system displays a high level of chromospheric activity, with strong and variable Mg~{\\sc ii} h+k and Ca~{\\sc ii} H+K emission (Hrivnak, Guinan, \\& Lu 1995). This suggests that OO Aql might also have variable chromospheric activity. In OO Aql, Ca~{\\sc ii} emission is not evident in medium-resolution spectra, although there may be infilling of the broad absorption profiles (Hrivnak 1989). However, this apparent absence is most probably the result of a contrast effect, since OO Aql is hotter and has a much higher continuum in this spectral region (3900$-$4000 {\\AA}) than does VZ Psc. Since the continuum is lower in the ultraviolet, the Mg~{\\sc ii} h+k emission at 2800 {\\AA} provides a better opportunity to measure the chromospheric activity in OO Aql. A single observation of OO Aql made with the low-dispersion LWR camera of the {\\it International Ultraviolet Explorer} ({\\it IUE}) satellite by Rucinski (1985) in 1984 indicated an unusually low Mg~{\\sc ii} emission level when compared with other W UMa-type systems. No x-ray detections, which would measure the coronal activity, have been reported for OO Aql, not even in the ROSAT Faint Source Catalogue (Voges et al. 2000). However, it has been measured as a variable radio source. It was one of only a few systems detected in a 3.6-cm radio continuum survey of W UMa-type binaries, but was not detected at similar sensitivity when observed two years later (Rucinski 1995). With this background of well-determined but unusual properties (compared with other W UMa-type binary systems), we were motivated to carry out an ultraviolet study of OO Aql with the {\\it IUE} satellite during a complete orbital cycle. Our goal was to investigate the level of its Mg~{\\sc ii} emission and to search for phase-related and time-related variations. As a by-product of these spectra, we also obtained ultraviolet light curves. In this paper we present the results of this study. ", "conclusions": "Ultraviolet observations were obtained for the short-period contact binary OO Aql on two dates in 1988, including nine consecutive observations that covered most of one orbital cycle. Mg~{\\sc ii} h+k emission is seen in each of the spectra. These were measured for emission strength, and also measurements were made of the continuum at several wavelength intervals and used to compose ultraviolet light curves for the binary. An analysis of the Mg~{\\sc ii} emission indicates that real variations in the strength of the feature exist for OO Aql, even when normalized to the surrounding ultraviolet continuum. These may be correlated with the orbital phase, or they may instead indicate overall changes in the chromospheric activity level with time. Previous studies of phase-related activity in W UMa-type binaries have yielded similar results. Eaton (1986) found only a slight phase dependence in the normalize Mg~{\\sc ii} emission for SW Lac (P=0$\\fd$32, K0), another contact binary with a relatively large mass ratio (0.73, Hrivnak 1992). The study of VZ Psc (Hrivnak et al. 1995) revealed variations in the normalized Mg~{\\sc ii} emission that were not phase dependent, but that seemed to vary in time. VW Cep (P=0$\\fd$27, G5V+G8V) shows primarily time-dependent variations in chromospheric Mg~{\\sc ii} emission and also in the transition region emission flux (Guinan \\& Gim\\'enez 1993). Thus variations in the normalized Mg~{\\sc ii} emission are common in W UMa-type binaries, but these are typically not phase dependent, and instead are probably due to variations in chromospheric activity arising from possible flaring. General trends have been found for W UMa-type and other short-period binaries in studies of chromospheric activity versus color or orbital period. A summary for W UMa-type binaries has been published by Rucinski (1985), which included his observation of OO Aql. While these more recent observations show OO Aql to have a somewhat higher level of chromospheric activity than at the time of his 1984 observation, the basic interpretation is the same; OO Aql has a much lower level of relative activity (by a factor of 2 relative to f(bol)) than do other contact binaries with similar colors. However, this is presumably due to the recent evolution of its components into contact. While in contact, OO Aql is expected to transfer mass from the less-massive secondary to the more-massive primary component (Webbink 1976), reducing the mass ratio and causing the system to evolve to evolve to an F spectral type. The level of chromospheric activity in OO Aql is in fact comparable to that in W UMa-type systems with the colors of F stars. Ultraviolet light curves are presented here for the first time for OO Aql. They show a similar shape and amplitude as the visible light curves. Since the two stars have similar temperatures, there is expected to be little change in the light curves with wavelength. Synthetic light curve solutions were carried out for $B$ and $V$ light curves for two different epochs, and similar orbital and physical parameters were obtained. These were used to model the new ultraviolet light curves, and good fits were obtained. The combination of ultraviolet and visible light curves led to a well-determined fit for a cool spot on the primary component. It appears that the light curves at the different epochs can each be fit by the same orbital and physical parameters, with the presence of one major cool spot on the primary component that varies in longitude and size with time. Jeong et al. (1994) carried out a somewhat similar ultraviolet and visible light curve analysis for SW Lac, also including a single cool spot. The high mass ratio and relatively long orbital period lead to a relatively large orbital angular momentum for OO Aql, which is larger than that of most contact binaries but less than that of the short-period detached binaries with G-type components, such as ER Vul and UV Leo. It is likely that the detached systems such as ER Vul (P=0$\\fd$70, M$_1$=0.96 M$_{\\sun}$, M$_2$=0.89 M$_{\\sun}$, Budding \\& Zeilik 1987) and UV Leo (P=0$\\fd$60, M$_1$=0.99 M$_{\\sun}$, M$_2$=0.92 M$_{\\sun}$, Popper 1980) will evolve into contact systems like OO Aql (P=0$\\fd$51, M$_1$=1.05 M$_{\\sun}$, M$_2$=0.88 M$_{\\sun}$) through angular momentum loss by magnetic breaking (Mochnacki 1981; Vilhu 1982). However, ER Vul has a much higher level of variable starspot and chromospheric activity than that seen in OO Aql, even though its rotational period is longer and its spectral type earlier than those of OO Aql. If OO Aql is typical of a system of solar mass stars that have recently evolved into contact, then the previous high level of activity must be significantly reduced in the contact phase. Subsequent evolution while in contact is expected to decrease the mass ratio, and if this is so, then a binary like OO Aql will gradually evolve from G to F spectral type as the mass is transferred from the secondary to the primary component. As shown by Rucinski (1985), the F spectral type systems show a lower activity, consistent with what is now seen in OO Aql. Thus this study of OO Aql has presented an opportunity to learn more about the properties of a close binary in a rather short-lived stage of its evolution, shortly after the two stars have evolved into contact. If OO Aql is typical of this stage, then it appears that the level of chromospheric activity must be reduced rather quickly upon entering the contact configuration, with the contact binary continuing to maintain this lower level with some smaller variations over time." }, "0011/astro-ph0011029_arXiv.txt": { "abstract": "The declining lightcurve of the optical afterglow of gamma-ray burst GRB000301C showed rapid variability with one particularly bright feature at about $t$$-t_0$=3.8 days. This event was interpreted as gravitational microlensing by Garnavich, Loeb \\& Stanek (2000) and subsequently used to derive constraints on the structure of the GRB optical afterglow. In this paper, we use these structural parameters to calculate the probability of such a microlensing event in a realistic scenario, where all compact objects in the universe are associated with observable galaxies. For GRB000301C at a redshift of $z$=2.04, the {\\sl a posteriori} probability for a microlensing event with an amplitude of $\\Delta m$$\\ge$0.95\\,mag (as observed) is 0.7\\% (2.7\\%) for the most plausible scenario of a flat $\\Lambda$-dominated FRW universe with $\\Omega_{\\rm m}$=0.3 and a fraction $f_*$=0.2 (1.0) of dark-matter in the form of compact objects. If we lower the magnification threshold to $\\Delta m$$\\ge$$0.10$\\,mag, the probabilities for microlensing events of GRB afterglows increase to 17\\% (57\\%). We emphasise that this low probability for a microlensing signature of almost a magnitude does {\\sl not} exclude that the observed event in the afterglow lightcurve of GRB000301C was caused by microlensing, especially in light of the fact that a galaxy was found within 2~arcsec from the GRB. In that case, however, a more robust upper limit on the {\\sl a posteriori} probability of $\\approx$5\\% is found. It does show, however, that it will not be easy to create a large sample of strong GRB afterglow microlensing events for statistical studies of their physical conditions on micro-arcsec scales. ", "introduction": "Gravitational microlensing offers a way to study the structure of high redshift sources on micro-arcsecond scales, besides being able to constrain the mass fraction and mass function of compact objects in the universe. In strong gravitational lens systems (i.e. systems with multiple images of a single background source), the microlensing optical depth is of order unity (e.g. Chang \\& Refsdal 1979; Gott 1981; Young 1981). Precisely because of this high optical depth, these systems are perfect for resolving micro-arcsec structure in the lensed cosmologically-distant source if it crosses a caustic created by the stellar-mass compact objects in the lens mass distribution (e.g. Chang \\& Refsdal 1984; Grieger, Kayser \\& Refsdal 1988; Wambsganss, Paczy\\'nski \\& Schneider 1990; Wo\\'zniak et al. 2000). Moreover, because of the presence of multiple images, one can in principle separate intrinsic source fluctuations from microlensing variability. For example, ongoing microlensing of the lensed (optical) images in the system Q2237+0305 (Huchra et al. 1985) has been observed ever since its discovery (e.g. Irwin et al. 1989; Corrigan et al. 1991; {\\O}stensen et al. 1996; Lewis et al. 1998; Wo{\\'z}niak et al. 2000). To a smaller degree and on longer time scales, microlensing in Q0957+561 has been detected as well (Pelt et al. 1998; see also Refsdal et al. 2000). The time scale of microlensing in both cases is defined by the relative transverse velocity between the source, lens and observer, which is given by the bulk velocity of the lensing galaxy plus the random motions of compact objects in the line-of-sight to the stationary quasar images, and is typically of order several hundred km\\,s$^{-1}$. This results in microlensing time scales of the order of months to years for solar-mass objects (e.g. Wambsganss 2000). \\begin{figure*} \\begin{center} \\leavevmode \\hbox{% \\epsfxsize=\\hsize \\epsffile{fig1r.eps}} \\end{center} \\caption{A cartoon of how we determine the microlensing magnification as a function of time for the expanding shell source superimposed on the magnification patterns caused by massive compact objects. The five panels indicate five different epochs (in practise, we consider 100 epochs). The superimposed curved line indicates the microlensing magnification as a function of time, corresponding to this particular example. The arrow points to a particular high magnification structure that is passed by the ring source at epoch $t_2$.} \\end{figure*} Besides these optical sources, recently the first case of radio-microlensing was reported in the lens system B1600+434 (Koopmans et al. 2000a, 2000b; Koopmans \\& de Bruyn 2000), suggesting extremely compact relativistic substructure in the lensed radio source. If this microarcsec-scale substructure is part of a relativistic jet, the time-scale of microlensing variability by stellar-mass objects reduces to several weeks (Koopmans \\& de Bruyn 2000), allowing one to probe compact objects up to $\\sim$$10^{5}$\\,M$_\\odot$ on time scales of several years, as well as the substructure of the relativistic jet. Indications of optical microlensing in B1600+434 with similar time scales have also been found (Burud et al. 2000). Clearly, microlensing is a promising field of future research regarding the study of high-$z$ sources at micro-arcsecond scales. In addition to testing the structure of AGNs, Loeb \\& Perna (1998) more recently proposed the use of microlensing to probe into the internal structure of GRB afterglows on micro-arcsec scales. Garnavich, Loeb \\& Stanek (2000) indeed interpreted an anomalous event in the lightcurves of the optical afterglow of GRB000301C (Masetti et al. 2000; Sagar et al. 2000; Berger et al. 2000; Jensen et al. 2000; Smette et al. 2000) as being caused by microlensing of the GRB afterglow. They subsequently derive constraints on its structure, which appear in good agreement with theoretical blastwave models. The inferred mass of the lensing object is $\\sim$0.5\\,M$_\\odot$, if its redshift is optimal for microlensing (i.e. about half way; Garnavich et al. 2000). However, one needs to be cautious here, because no multiple images are present -- as in the case of strong gravitational lenses -- to confirm that this is indeed a non-intrinsic event. The fact that the event occurs within only a few days after the burst and has an amplitude of $\\approx$1 mag, suggests that the burst must have occurred close to the Einstein radius of an intervening massive compact object (Garnavich et al. 2000). To have a significant probability of observing such a GRB microlensing event, the universe requires a surface density in compact objects close to the critical surface density (e.g. Press \\& Gunn 1973; Blaes \\& Webster 1992). The situation is partly similar to the case of variability in single quasars (i.e. not multiply-imaged), which also cannot easily be proven to be due to microlensing (e.g. Hawkins \\& Taylor 1997; Hawkins 1998), although the freedom to model the GRB afterglow lightcurve (mostly dominated by self-similar expansion resulting in a power-law behavior) is significantly less than that for quasars. In this paper, we investigate the probability of microlensing in GRB afterglows for a flat $\\Lambda$--dominated cosmological model and with the distribution of massive compact objects connected to visible galaxies, particularly focusing on the strong event seen in the GRB000301C afterglow. Our conclusions, however, are independent of this particular GRB. In Section 2, we describe numerical microlensing simulations of GRB afterglow light curves, extended to high microlensing optical depths. In Section 3, the {\\sl a posteriori} probability of the observed event in GRB000301C is calculated. Section 4 summarizes our result and states our conclusions. ", "conclusions": "If the event seen in the optical lightcurve of the GRB000301C afterglow is caused by microlensing, the opportunities to study the structure and evolution of GRB afterglows on micro-arcsec scale are potentially very exciting (Loeb \\& Perna 1998; Garnavich et al. 2000; Mao \\& Loeb 2000). However, in this paper we have shown that the {\\sl a posteriori} probability of this particular event is actually small, i.e. 0.7--2.7\\% for a fraction $f_*$=0.2--1.0, respectively, of dark-matter in the form of compact objects. The main assumptions in this calculation are: (i) a constant comoving density population of galaxies which follow a Schechter luminosity (i.e. mass) function, (ii) a flat $\\Lambda$--dominated FRW universe with $\\Omega_{\\rm m}$=0.3, (iii) all matter in the universe traces galaxies, which can be described as singular isothermal spheres, (iv) the mass spectrum of compact objects is `narrow', in which case they can be parameterised by a single value for their mass, (v) a typical GRB optical afterglow has similar properties as GRB000301C and has a redshift of $z$$\\approx$2, and (vi) there is no significant magnification bias. The magnification bias, however, is unlikely to be a problem. Even in case this bias increases the number of observed GRBs by a factor of 10 for high microlensing optical depths ($\\kappa$$\\ga$0.25), Table~1 and eqn.(7) show that the overall probabilities are increased by only 10$\\times P(\\Delta m \\ge 0.95^{\\rm m}|\\kappa\\ga0.25) \\times\\tau_{\\rm GL/GRB}$$\\sim$0.1\\%. A detailed calculation should consistently take into account both the magnification distribution of galaxies {\\sl and} that of compact objects (e.g. Pei 1993a, 1993b). In light of the very uncertain redshift distribution and luminosity function of GRB afterglows this is, however, not yet warranted. Moreover, even in the case that the {\\sl whole} sky has a surface mass density of $\\kappa_*$$\\la$0.25 (normalised by the ``critical'' surface mass density for lensing, e.g. Schneider et al. 1992) our simulations (see Table~1) indicate that the probability of such an event does not significantly exceed $\\approx$5\\%. Clearly, the latter situation is unrealistic, but it does place a very robust upper limit on the probability of this event (see also Press \\& Gunn 1973 and Blaes \\& Webster 1992). The plausibility to actually observe such a fraction of GRB afterglow lightcurves with a microlensing event of about one magnitude also depends on the typical mass of compact objects. However, even if the mass goes up or down by a factor of ten from the assumed 0.5\\,M$_\\odot$, to first approximation that would just mean a ``broadening'' in time of the microlensing event by about a factor of three (since the length/time scale is proportional to the square root of the mass). However, if most compact objects have masses {\\it much} smaller than that inferred for GRB000301C, the fraction of GRB afterglows with microlensing signatures will obviously decrease. In that case the effect of the ``shrunk'' caustics will average out over the much larger angular size of the emission region of the GRB afterglow. This case, however, seems unlikely in light of the stringent lower limit ($\\sim$10$^{-2}$\\,M$_\\odot$) placed on low-mass compact objects in 0957+561 (Schmidt \\& Wambsganss 1998; Refsdal et al. 2000; Wambsganss et al. 2000). The fraction of microlensed GRB afterglows would also decrease if the typical redshift of GRBs is significantly lower than~$z=2$. A somewhat lower typical redshift is supported by the average redshift $_{\\rm GRB}\\approx1.3$, with an rms spread of $\\approx$1, that we find from Bloom, Kulkarni, \\& Djorgovski (2000). In case of a broad mass spectrum of the compact objects or a large spread in the physical properties of the GRB optical afterglows, one has to convolve the microlensing probabilities with these properties. This, however, is not expected to change the results significantly. We emphasise that our results {\\sl do not exclude} that the event seen in the lightcurve of the optical afterglow of GRB000301C is in fact due to gravitational microlensing, especially in light of the fact that a galaxy was found at 2~arcsec from the GRB (Garnavich et al. 2000). Each event should be treated on its own merit (and it is a posteriori anyway). In any case, in the sample of $\\sim$20 known GRBs with observed optical afterglows (e.g. Bloom et al. 2000), the probability of one such event would be around 14\\%. But our results do predict that it is unlikely to find many strongly microlensed GRB afterglow soon, and it emphasises the difficulties one will encounter in creating a large sample of these strong (i.e. $\\Delta m \\ga1$\\,mag) events, with the aim to statistically study the physical properties of GRB afterglows on micro-arcsec scales. For example, to get 10 additional microlensing events of comparable strength, one would require $\\approx$1500 GRB afterglow optical lightcurves. Even for the {\\sl Swift} satellite (Parsons et al. 1999; see also Mao \\& Loeb 2000) this provides a challenging task. On the bright sight, if the event in GRB000301C was due to microlensing, we can expect a significant fraction of high-$z$ GRB afterglow light curves to show microlensing events stronger than 0.1\\,mag (see also Mao \\& Loeb 2000), although these will be hard to distinguish from intrinsic variations and will also not really be able probe the GRB afterglow structure on micro-arcsec scales. We conclude, that the best cases to unambiguously study micro-arcsec structure in GRB afterglows will probably be those GRBs that are multiply imaged (about 1--2\\% expected, see, e.g. Holz et al. 1999; Komberg et al. 1999; Mao 1992, 1993; Marani et al. 1999; McBreen et al. 1993; Narayan \\& Wallington 1992; Nowak \\& Grossman 1993; Paczy\\'nski 1987; Wambsganss 1993; Williams \\& Wijers 1997). Those (soon to be discovered!) will almost certainly (see Table~1) show microlensing events $\\ga$0.1\\,mag, which can easily be separated from intrinsic variations by comparison with the other lensed GRB images, after correcting for the respective time delays. A smaller fraction ($\\approx$5\\%) will also show events $\\ga$1\\,mag, although this might increase due to the magnification bias. So the prospects of doing science with (micro-)lensed gamma ray bursts and afterglows remain promising." }, "0011/astro-ph0011503_arXiv.txt": { "abstract": "Hydrodynamical winds from a spherical two-temperature plasma surrounding a compact object are constructed. The mass-loss rate is computed as a function of electron temperature, optical depth and luminosity of the sphere, the values of which can be constrained by the fitting of the spectral energy distributions for known X-ray binary systems. The sensitive dependence of the mass loss rate with these parameters leads to the identification of two distinct regions in the parameter space separating wind-dominated from non wind dominated systems. A critical optical depth ($\\tau_c$), as a function of luminosity and electron temperature, is defined which differentiates these two regions. Systems with optical depths significantly smaller than $\\tau_c$ are wind-dominated. The results are applied to black hole candidate X-ray binary systems in the hard spectral state (Cyg X-1, GX 339-4 and Nova Muscae), and it is found that the inferred optical depth ($\\tau$) is similar to $\\tau_c$ suggesting that they are wind regulated systems. On the other hand, for X-ray binary systems containing a neutron star (e.g., Cyg X-2) $\\tau$ is much larger than $\\tau_c$ indicating the absence of significant hydrodynamical winds. ", "introduction": "\\label{sect1} X-ray binary systems containing a black hole candidate are typically found to be in two different spectral states. In the hard state, the broad band X-ray spectrum can be described as a power-law (photon spectra index $\\approx 1.5$) with a high energy cutoff around 100 keV. In the soft state, the spectrum consists of two components. There is usually an extended power-law (with spectral index $\\approx 2.5$) and a soft X-ray emission which has a spectral shape similar to that of a black body. For a recent review of the observations and phenomenological description of these sources see Tanaka \\& Shibazaki (1996). The modeling of the hard state spectra can be described in terms of an unsaturated Comptonization process of soft photons in a region with hot electrons ($kT \\approx 50$ keV) and electron scattering optical depths of order unity. In a pioneering study, Shapiro, Lightman \\& Eardley (1976) identified this hot region with a geometrically thick, optically thin, hot accretion disc. In this model, the gravitational energy dissipated in the disc, heats the ions which in turn transfer the energy to electrons by Coulomb interactions. However, the electron-ion Coulomb interaction rate is inefficient in such an environment of low density and high electron temperature, and this leads to a large difference in temperature between the electrons and ions, with the ion temperature reaching nearly virial values ($10^{11} K$). The importance of radial advection of energy in such a disc was noted by Ichimaru (1977). Taking advection into account Narayan \\& Yi (1994) constructed self-similar solutions for the disc equations called Advection Dominated Accreting Flows (ADAF) showing not only that the proton temperatures approach their virial values, but also that the radiative efficiency of accretion can be significantly reduced as a result of the advection of energy into the black hole. In an alternative description Chakrabarti \\& Titarchuk (1995) argued that under certain conditions a shock may arise in such accretion discs and identified the hot Comptonizing region with the post-shock flow. Despite the differences in the geometry, radiative processes and the detailed disc structure, both these models have in common the presence of a two-temperature plasma. Such a plasma is a natural outcome of any accretion disc model which (a) identifies the hard state X-ray spectrum as a result of the unsaturated Comptonization process of soft photons, (b) assumes that the viscous energy dissipated heats the ions and (c) that the only mechanism for energy transfer between the ions and electrons is Coulomb interaction. We note here that these assumptions may not be valid since the viscous energy dissipated may heat the electrons preferentially if a strong equipartition magnetic field is present in the disc (Bisnovatyi-Kogan \\& Lovelace 2000). In addition, there could be unknown mechanisms which transfer energy between ions and electrons more efficiently than the Coulomb interaction. Thus, it will be useful to have an independent observational signature which could confirm the existence of two-temperature plasmas in black hole candidate systems. The nearly virial proton temperature of this plasma suggests the possibility of a strong hydrodynamical wind arising from these systems (e.g., Piran 1977; Takahara, Rosner \\& Kusunose 1989; Kusonose 1991). Such an outflow could transport away a significant fraction of mass, energy and/or angular momentum, thereby affecting the structure and stability of the disc and affecting the radiative efficiency of accretion for a given mass transfer rate. Chakrabarti (1999) and Das (1999) have studied the possibility of outflows in the context of the shock/centrifugal barrier models. They find that for certain values of the disc parameters (e.g., accretion rate and specific entropy) a hydrodynamical wind occurs. For the ADAF disc solutions, Blandford \\& Begelman (1999) argued that only a small fraction ($<$ 1\\%) of the gas actually falls into the black hole and the rest is driven away as a wind (in an advection dominated inflow-outflow solution - ADIOS), thereby effectively reducing the radiative efficiency. Beckert (2000) confirmed this result for different viscosity laws while Quataert \\& Narayan (1999) showed that the X-ray spectra from such a wind driven accretion process can explain the observed spectra of some black hole candidate systems in quiescence. These calculations were undertaken for low mass accretion rates, and it is not clear how the system will behave in the high accretion rate regime inferred for black hole candidate systems in the hard state. The formation of hydrodynamical winds from accretion discs depends on the structure and geometry of the discs. Thus, detailed calculations of the outflow are intrinsically model dependent. Further, reliable calculations of the structure of such discs are difficult because uncertainties exist in the vertical distribution of energy dissipation in the disc associated with our lack of detailed understanding of the viscosity. We note that internal magnetic fields in the disc could also facilitate (or inhibit) the formation of winds. Here, electromagnetic forces may accelerate and collimate the wind to form high velocity jets as observed in microquasar systems (Mirabel \\& Rodriguez 1999). Considering these uncertainties a prudent approach would be to estimate the wind characteristics using only those parameters which can be directly constrained by the fitting of spectral energy distributions. Such an analysis would allow a rough estimation of the magnitude of the mass and energy lost in the form of a wind for a system in question. With such an objective in mind, we report in this paper on calculations of the structure of an hydrodynamical wind similar to those found by Takahara et al (1989), but in the context of a uniform spherical two-temperature plasma around an accreting black hole. Rather than treating the entire disc/wind configuration with detailed heating and cooling processes as in Kusunose (1991) the calculations are parameterized in terms of the electron temperature of the cloud, $T_e$, the optical depth, $\\tau$, and the total luminosity of the source. We adopt this approach as these parameters are constrained by spectral fitting analyses. In the next section, the formulation of the problem and the numerical results are presented. The application of these results to observed systems is given in \\S 3 and discussed in the last section. ", "conclusions": "\\label{sect4} The possible occurrence of hydrodynamical winds in X-ray binary systems has been investigated. From simple considerations, the conditions under which these winds can be important have been identified. It is found that the mass loss rate in such winds depends sensitively on the luminosity of the source as well as the electron temperature and optical depth of the coronal region. The steep dependence of the mass loss rate on these parameters facilitates the use of a critical optical depth to indicate whether a given system can support such a wind. Application of the theory indicates that winds can exist in the hard state of the black hole candidates Cyg X-1, GX 339-4 and GS 1124-68. Strong winds in these systems may decrease the radiative efficiency ($\\eta = L/\\dot M_i c^2$, where $\\dot M_i$ is the mass inflow rate) by carrying away a substantial amount of matter and energy. For Nova Muscae, the radiative efficiency has been estimated to be $\\eta = 0.01$ for the soft state and $\\eta = 0.05$ for the hard state (Misra 1999). These values are lower than that expected from an Keplerian disc ( $\\eta \\approx 0.1$) which could be due to the presence of a strong wind rather than energy advection into the black hole. Winds may also effect the thermal stability of accretion discs by introducing an additional channel for energy loss (e.g., Piran 1977). We reiterate that the analysis undertaken in this paper is based on the assumption that the gravitational energy dissipated in the system is transferred to the electrons by the ions via Coulomb interactions. This naturally restricts the analysis to only certain systems where this is valid. As mentioned earlier, the soft state spectrum of black hole candidate systems is probably of a non-thermal origin and, hence, the analysis performed in this study cannot be applied. Even for the hard spectral states of compact X-ray binaries considered here, alternate models to a two -temperature plasma description have been proposed. For example, the X-ray spectra could be due to magnetic flare activity above a cold disc (Poutanen \\& Fabian 1999) or produced by a disc with a rapidly varying radial temperature profile (Misra, Chitnis \\& Melia 1998). Thus, it should be emphasized that the results presented here are specifically discussed within the framework of a two-temperature plasma model. The spherical geometry assumed here is simplistic even though two-temperature discs are geometrically thick. Further the effect of angular momentum or of convection on the dynamics of accretion and the wind outflow have not been taken into account. The former effect can increase the mass outflow rate as a result of centrifugal support, whereas the latter effect can affect the thermal structure of the underlying disk and, hence, the existence of a wind. In this context, the thermal structure is dependent on the direction in which angular momentum is radially transported by convection and the magnitude of the viscosity parameter (see Narayan, Igumenshchev, \\& Abramowicz 2000). A more detailed analysis should take these effects and radiative heating/cooling of the wind into account in at least two spatial dimensions (e.g., Chakrabarti \\& Molteni 1993). However, such analyses will, by necessity, be model dependent and limited by the uncertainties in the vertical structure and geometry of the hot disc. The hydrodynamical winds described in this paper may be confined to form a jet-like structure by the geometry of the disc and magnetic fields. Further if they are accelerated to relativistic speeds by electromagnetic forces, they may provide the origin of the radio jets observed in the black hole candidate systems known as microquasars (Mirabel \\& Rodriguez 1998). On the other hand, the radio jets may be a different phenomenon unrelated to hydrodynamical winds. In that case it is desirable to have direct observational signatures of these outflows. Since the Thomson optical depth of the winds calculated here (i.e. the optical depth from the surface of the sphere $R$ to infinity) is typically $< 0.1$, column densities of the order of $10^{23}$ cm$^{-2}$ are indicated. Although the X-ray continuum spectra are not expected to be altered by the outflow, a significant fraction of the column density may not be highly ionised giving rise to observable absorption and/or emission lines. The detection of P-Cygni type profiles by high resolution X-ray satellites Chandra and Newton-XMM may provide evidence for the existence of such winds and serve as a useful diagnostic not only of the wind structure, but also the physics underlying the disc as well. \\bigskip RM acknowledges support from the Lindheimer Fellowship at Northwestern University." }, "0011/astro-ph0011359_arXiv.txt": { "abstract": "We discuss the ultraviolet to near-IR galaxy counts from the deepest imaging surveys, including the northern and southern {\\it Hubble Deep Fields}. The logarithmic slope of the galaxy number-magnitude relation is flatter than $0.4$ in all seven $UBVIJHK$ optical bandpasses at faint magnitudes, i.e. the light from resolved galaxies has converged from the UV to the near-IR. Most of the galaxy contribution to the extragalactic background light (EBL) comes from relatively bright, low-redshift objects (50\\% at $V_{\\rm AB}\\lta 21$ and 90\\% at $V_{\\rm AB}\\lta 25.5$). We find a lower limit to the surface brightness of the optical EBL of about $15\\,\\eblunits$, comparable to the intensity of the far-IR background from {\\it COBE} data. Diffuse light, lost because of surface brightness selection effects, may add substantially to the EBL. ", "introduction": "The extragalactic background light (EBL) is an indicator of the total luminosity of cosmic structures, as the cumulative emission from pregalactic, protogalactic, and evolving galactic systems, together with active galactic nuclei (AGNs), is expected to be recorded in this background. The recent progress in our understanding of faint galaxy data, made possible by the combination of {\\it Hubble Space Telescope} {(\\it HST}) deep imaging and ground-based spectroscopy, and of the evolution of the stellar birthrate in optically-selected galaxies from the present-epoch up to $z\\approx 4$ (Steidel \\etal 1999; Madau, Pozzetti, \\& Dickinson 1998), has been complemented by measurements of the far-IR/sub-mm background by the {\\it COBE} satellite (Hauser \\etal 1998; Fixsen \\etal 1998; Puget \\etal 1996), showing that a significant fraction of the energy released by stellar nucleosynthesis is re-emitted as thermal radiation by dust (Dwek \\etal 1998). In this talk I will focus on the galaxy number-apparent magnitude relation and its first moment, the galaxy contribution to the EBL. The logarithmic slope of the differential galaxy counts ($d\\log N/dm_{\\rm AB}\\equiv \\gamma$) is a remarkably simple cosmological probe of the history of the stellar birthrate, as it must drop below 0.4 to yield a finite value for the EBL. The radiation emitted from unresolved sources that could be lost due to uncertainties in the faintest galaxy data and surface brightness selection effects will be discussed, together with the contribution to the EBL from high-$z$ populations such as the Lyman-break galaxies and extremely red objects. \\begin{figure} \\centerline{\\psfig{figure=EBL_diff.ps,width=8cm}} \\caption{\\small Extragalactic background light per magnitude bin, $i_\\nu=10^{-0.4(m_{\\rm AB}+48.6)}N(m)$, as a function of $U$ ({\\it filled circles}), $B$ ({\\it open circles}), $V$ ({\\it filled pentagons}), $I$ ({\\it open squares}), $J$ ({\\it filled triangles}), $H$ ({\\it open triangles}), and $K$ ({\\it filled squares}) magnitudes. For clarity, the $BVIJHK$ measurements have been multiplied by a factor of 2, 6, 15, 50, 150, and 600, respectively.} \\end{figure} \\begin{figure} \\centerline{\\psfig{figure=EBL_cum.ps,width=7cm}} \\caption{\\small Cumulate contribution to the EBL per magnitude bin as a function of $U,B,V,I,J,H,K$ AB magnitudes, derived from a fit to the observed counts.} \\end{figure} ", "conclusions": "" }, "0011/astro-ph0011390_arXiv.txt": { "abstract": "s{ We present the results of a redshift survey of both cluster and intercluster galaxies in the central part of the Shapley Concentration, the richest supercluster of clusters in the nearby Universe, consisting of $\\sim 2000$ radial velocities. We estimate the total overdensity in galaxies of the supercluster and its mass and we discuss the cosmological implications of these results. Moreover, using a Principal Components Analysis technique, we study the influence of the cluster and supercluster dynamics on the galaxy spectral morphology. } \\noindent Clusters of galaxies are thought to form by accretion of subunits in a hierarchical bottom up scenario: numerical simulations revealed that mergings happen along preferential directions, called density caustics, which define matter flows, at whose intersection rich clusters are formed. \\\\ The central regions of rich superclusters, where the overdensity is of the order of $\\sim 10$, are thought to be in the early collapse phase and therefore give the possibility to study the formation of clusters and groups of galaxies. These regions are the ideal environment for the detection of cluster mergings, because the peculiar velocities induced by the enhanced local density of the large scale structure favour the cluster-cluster and cluster-group collisions, in the same way as the caustics seen in the simulations. \\\\ Spectacular major merging events are visible in the central region of the Shapley Concentration, the richest supercluster of galaxy clusters in the nearby Universe (Zucca et al. 1993). Particularly interesting are two chains of interacting clusters: the A3558 and the A3528 complexes (Fig. 1). \\\\ We are carrying on a long term multiwavelength study on the Shapley Concentration (see also the poster by Bardelli et al., these proceedings), in order to determine its mass, dynamical state and cosmological relevance, as well as to analyse in detail the physics of the merging phenomena. \\\\ In this poster we focus on the large-scale distribution and spectral properties of galaxies both in clusters and in the ``field\" in between clusters. \\begin{figure} \\begin{center} \\epsfxsize=0.7\\hsize \\rotate[r] { \\epsfbox {fig1a_iap.ps} } \\epsfxsize=0.7\\hsize \\rotate[r] { \\epsfbox {fig1b_iap.ps} } \\end{center} Figure 1: Optical distribution of galaxies in the central part of the Shapley Concentration. \\end{figure} ", "introduction": " ", "conclusions": "" }, "0011/astro-ph0011445_arXiv.txt": { "abstract": "The time evolution of initially balanced, rapidly rotating models for an isolated disk of highly flattened galaxies of stars is calculated. The method of direct integration of the Newtonian equations of motion of stars over a time span of many galactic rotations is applied. Use of modern concurrent supercomputers has enabled us to make long simulation runs using a sufficiently large number of particles $N=30,000$. One of the goals of the present simulation is to test the validities of a modified local criterion for stability of Jeans-type gravity perturbations (e.g. those produced by a barlike structure, a spontaneous perturbation and/or a companion galaxy) in a self-gravitating, infinitesimally thin and collisionless disk. In addition to the local criterion we are interested in how model particles diffuse in velocity. This is of considerable interest in the kinetic theory of stellar disks. Certain astronomical implications of the simulations to actual disk-shaped (i.e. rapidly rotating) galaxies are explored. The weakly nonlinear, or quasi-linear kinetic theory of the Jeans instability in disk galaxies of stars is described as well. ", "introduction": "In galactic astronomy, the most intensely studied of the problems of pattern formation is the problem of the formation of spiral structure, which has brought forth numerous theoretical and numerical approaches. Even though at the present time there exists, as yet, no satisfactory theory of the origin and conservation of the spiral structure so prominent in the Milky Way and many other giant flattened galaxies, the majority of experts in the field has yielded to the opinion that the study of the stability of the small-amplitude cooperative oscillations in disk-shaped galaxies of stars is the first step towards an understanding of the phenomena. This is because in the Milky Way and many other disk galaxies, the bulk of the luminous mass, probably $\\gtrsim 90\\%$, is in the stars. Therefore, it seems likely that the spiral structures are intimately connected with the stellar constituent of a galaxy, and that stellar dynamical phenomena play a basic role. Because of the long-range nature of the gravitational force between stars, a galaxy exhibits collective modes of motions --- modes in which the stars in large regions move coherently or in unison. A galaxy is characterized also by a high ``temperature\" (a dispersion of random velocities of stars) and hence a high ``thermal\" energy; this thermal energy is much larger than the interaction potential between pairs of stars. Because of this, binary encounters produce only small-scale scattering of the motions of the stars. Over long enough times, these gradually build up to produce large deflections that constitute collisions. In many cases these collisional effects are so weak that we consider the galaxy as collisionless on the Hubble time $T \\sim 10^{10}$ yr (Chandrasekhar, 1960; Binney and Tremaine, 1987). Thus, the galaxy is dominated by the collective motions and the free streaming of the stars (kinetic effects). As usual it is very difficult determine a priori whether a particular linearization of nonlinear equations made in analytical studies constitutes a valid approximation. This can be determined, however, by direct computer simulations of the nonlinear theory that mimic the behaviour of stars in galaxies. Such modeling gives more detailed information than can be obtained analytically, so that the important physical phenomena can be determined. In this paper, the results of the approximate theoretical analysis described in the Appendix are compared with our own many-body ($N$-body) simulations. One of the goals of the simulation is to test the validities of a modified local criterion for stability of Jeans-type gravity perturbations in a self-gravitating, infinitesimally thin and collisionless disk. The fact that the nonaxisymmetric perturbations in the differentially rotating system are more unstable than the axisymmetric ones is taken into account in this modified Toomre-like (Toomre, 1964, 1977) criterion. In addition to the stability criterion we are interested in how model particles diffuse in velocity. As for the present study, such random velocity diffusion is caused by stellar disk turbulence. In particular, we verify the analytical prediction that as the Jeans-unstable perturbations grow the mean-square velocity increases linearly with time. The observation of this behavior is a sensitive test of the model as well as the kinetic theory of stellar disks. In fact, Morozov (1981) and Griv {\\it et al.} (1999a) already attempted to confirm the modified criterion numerically. However, because of the very small number of model stars, $N=200$, Morozov's results are subject to considerable uncertainties, and additional simulations are clearly required to settle the issue. Furthermore, for that number of model particles, the two-body relaxation time scale is comparable to the crossing time, even with Morozov's modest softening parameter, raising some question about the applicability of his simulations to actual almost collisionless galaxies. Increasing the number of model stars is definitely a more reliable procedure. In turn, Griv {\\it et al.} used a different numerical approach of the so-called local $N$-body simulations. The $N$-body experiments in a local or Hill's approximation has been pioneered by Toomre (1990), Toomre and Kalnajs (1991) and Salo (1995). In these simulations dynamics of particles in small regions of the disk are assumed to be statistically independent of dynamics of particles in other regions. The local numerical model thus simulates only a small part of the system and more distant parts are represented as copies of the simulated region. Unlike Morozov (1981) and Griv {\\it et al.} (1999a), we study some aspects of dynamical behavior of stellar systems by global simulations using a sufficiently large number of particles. ", "conclusions": "The results show that a velocity dispersion given by Toomre's (1964) criterion will stabilize a disk only against axisymmetric ring-like gravity perturbations. However, such disks are unstable against nonaxisymmetric spiral-like perturbations. Jeans-unstable spiral perturbations ``heat\" the system; the mean-square random velocity increases linearly with time. The results are in agreement with analytical predictions. We were able to generate an axisymmetric, gravitationally stable disk. The initial condition for the axisymmetric stable disk was obtained analytically and confirmed experimentally: the radial dispersion of random velocities of stars should be equal (or greater than) to the modified dispersion at each radii." }, "0011/hep-ph0011310_arXiv.txt": { "abstract": " ", "introduction": "Many ultra high energy (UHE) cosmic ray air showers with energies in excess of $5\\times 10^{19}$ eV have been observed in the past few decades \\cite{nw}. The nature and origin of the primary particles is not understood \\cite{nw,es,sigl}. The puzzle is that the sources have to be within the GZK limit of approximately 50 Mpc if these particles are protons or nuclei \\cite{GZK,puget}. However there are not enough powerful astrophysical sources within this distance to explain the events. Among known particles, only neutrinos travel larger distances than protons in intergalactic space. This leaves neutrinos as the only established candidates that can travel the distances greater than 100 Mpc from known UHE sources. The GZK bound of 50 Mpc is not applicable to them. Yet neutrino interactions with matter are too weak in the Standard Model of particle physics to generate the observed air showers. Hence these events seem to demand a revision of our current understanding of nature. Either the determination of the number of sources of such ultrahigh energy particles in our astrophysical neighborhood is grossly low\\footnote{For example if magnetic fields outside the galaxy have been underestimated, ``line of sight'' and ``photon travel time'' requirements on protons and nuclei can be relaxed and new source possibilities considered \\cite{es}}, or these observations are a signal of new physics. Many speculative ideas have been proposed to explain the events above $10^{19} - 10^{20}eV$, including topological defects such as cosmic strings \\cite{abs,mb} and associated decays of heavy, relic particles \\cite{berez,kt}, existence of neutral, stable, strongly interacting particles, such as a light gluino \\cite{ch,raby,lou} or a monopole bound state \\cite{kw,wick}, and violation of Lorentz invariance \\cite{g-m,cg}. Much of this work requires that the primary particle responsible for generating these air showers is an exotic new particle which does not exist within the Standard Model. In a recent paper \\cite{us1} we argued that the data is consistent with the general features of massive spin-2 exchange. Models where effects of gravity can be strong just above the weak scale \\cite{add} supply a natural and attractive framework. The interaction cross section of neutrinos with matter is greatly enhanced with massive spin-2 exchange at UHE and may reach values close to the hadronic cross sections. In the low scale gravity models, the cross section enhancement arises from t-channel exchange of the tower of gravitons. Our estimates of the neutrino-proton cross section at the highest energies relevant for these events are of the order of one to a few hundred millibarns. The highest energy cosmic ray events may therefore be initiated by neutrinos\\footnote{Correlation between the positions of compact radio quasars and the track directions of UHE $>100$ EeV cosmic rays has been studied by several groups \\cite{fb,sigl2,us2}.}\\cite{us1,early}. A generic, robust prediction of massive spin-2 exchange, known for many decades, is that the total cross section should grow with a power of energy, typically $\\sigma_{tot} \\sim s^2$. The property of power law growth with a power exceeding 1 (the result of 4-Fermi {\\it spin-1} exchange) is quite hard to evade and can be traced to dimensional analysis. We consider large cross sections at $UHE$ to be characteristic of extra-dimension, low scale gravity models. Interaction of $UHE$ neutrinos is a quite natural domain to seek the new effects of low scale gravity models: the very weakness of the Standard Model neutrino coupling minimizes this background, while the regime of of highest possible energy maximizes the effects of graviton-KK mode exchange. The theory of low-scale gravity models is only partly developed, and questions of unitarity complicate the interpretation of perturbation theory \\cite{gw}. One can choose models of the cross section which are further from the calculations of perturbation theory in the sense that they grow at a slower rate with energy than $s^2$ (The perturbative, parton level cross section rises as $\\hat s^3$) \\cite{N+S,kp,usnote}, or which operate by a separate (s-channel) mechanism \\cite{d+d}. Indeed it is possible to restrict models of low-scale gravity to the extent that nothing observable is predicted at the energies in question. For example, the astrophysical bound on the scale parameter $M$ for the n=2 case guarantees that the consequences of this model are unobservable \\cite{N+S},\\cite{kp}. There has been some confusion on this point. Let us compare the model we use here, taken from our previous cross section calculations \\cite{us1}, with subsequent work \\cite{kp}. The latter reports the result of assuming that a finite brane tension introduces an exponential damping of higher KK modes, providing an alternative cutoff mechanism \\cite{bando}. Like our calculation, when $\\surd s \\geq M$, the cross section in \\cite{kp} rises approximately quadratically with neutrino energy (See Fig. 1 in \\cite{kp}, where $\\sigma_{\\nu N}$ rises by two orders of magnitude for every order of magnitude rise in $E_{\\nu}$). Unlike ours, the calculation there assumes n=2 only, for which SN1987a analysis makes the restriction $M$ $\\geq$ 30-70 TeV \\cite{bounds}.\\footnote {Citing uncertainties in astrophysical parameters, \\cite{kp} considers $M$ values as low as 6 TeV.} If $n \\geq 3$ were considered, the scale could be lowered to the 2-3 TeV range and cross section values in agreement with ours would result. This is clear from the trend with mass scale in Fig. 1 in \\cite{kp}. Conversely, we could suppress our cross section to their values by raising $M$ to values of 6 TeV and above. Specifically, we find that $\\beta = 1$ and $M$ = 6 TeV yields $\\sigma_{\\nu N}$ = 0.3 mb, compared to 0.1 mb at $E_{\\nu}$ = $10^{20}$ TeV for $M$ = 6 TeV in \\cite{kp}, while the choice $\\beta$ = 2 and $M$ = 6.6 TeV or $\\beta$ = 1 and $M$ = 7.3 TeV reproduces their 0.1 mb value. Within modest parameter variations, the results clearly agree. This is not a surprise, since the parton level amplitudes and cross sections for small t are essentially identical, and are insensitive to the value of n in the two cutoff methods \\cite{bando}. The two calculations differ only in the details of the large t cutoffs, both of which produce $s^{2}$ behavior of the cross sections. The cutoff used in in \\cite{kp} gives a cross section result in essential agreement with ours at a given set of $E_{\\nu}$ and $M$ values. Their assertion to the contrary is an unfortunate consequence of presenting the results for a lower bound $M \\geq 6$, justified only for n=2, and drawing sweeping, unjustified conclusions about the general situation. With this technical issue clarified about the particular model we employ, we reiterate that our goal is to explore the broad consequences of strongly interacting $UHE$ neutrinos. The primary, general question that arises, defining our focus here, is the nature of {\\it air shower development}. High-energy leptonic interactions do not have the same multiplicity or inelasticity as high energy hadronic interactions. To understand the potential relevance of neutrino interactions, one must address not only the total cross section, but also the way the interaction delivers energy into the air showers that are actually observed. In the present paper we compare simulations of air showers generated by neutrino primaries with large cross sections to those generated by protons in the Standard Model. We ask whether there is anything about existing showers which might {\\it rule out} neutrinos as primaries.\\footnote{Alternatively one might ask whether one can ``find evidence'' for neutrinos as the primaries in some features of the showers. We do not pursue this, because the fluctuations of air showers and flexibility in simulation codes make it a very hard and ambiguous way to proceed.} If so then the case is made that large cross sections alone are not enough to support the case for neutrinos, and speculative models of \\emph{new} particles might be indicated. Contrary to some expectations \\cite{kp}, we find that neutrinos with large cross sections {\\it can create showers that are much like proton-initiated showers} and that in some cases are indistinguishable from them \\cite{sm} \\cite{ddm}. Two features of the low-scale gravity contribution to the neutrino-nucleon cross section come into play: first, the cross section is large enough to initiate air showers at high enough altitudes; second, its rapid $s^2$ dependence suppresses new effects among secondary products, which carry at most a few percent of the primary energy. Our methodology can evidently be extended to other models for hadronic size UHE neutrino cross sections provided the cross sections grow rapidly (as in spin-2 exchange). Other speculative primaries should be considered on a case-by- case basis. One cannot take the existence of one model that produces well simulated showers above $10^{20}$ eV to be conclusive evidence for a given hypothesis for the identity of the primary, neutrino or otherwise. The question of the mysterious primary, then, needs to be framed in view of everything that can be observed: cross sections, shower characteristics, and angular distributions and correlations, which may be informative about the charge of the primary. ", "conclusions": "The conjecture that UHE, super-GZK cosmic ray showers are caused by neutrinos with large UHE cross sections is almost as old as the field itself \\cite{early}. Revisiting this idea in a new theoretical framework \\cite{add}, we proposed models that achieve interestingly large cross sections \\cite{us1}. We speculated that the GZK ``barrier'' could be then broken by neutrinos. The next question to answer is whether the shower events predicted look like events observed. Or are the characteristics so different from observed showers, which are generally compared to those of the simulations of proton, nucleus and gamma initiated showers, that the neutrino can be eliminated as a candidate for the super GZK showers? {\\it Our conclusion based on this study is no, they cannot be eliminated as candidates}. For a range of values of the fundamental scale $M$ in the neighborhood of $2 -3$ TeV, there are neutrino energies 25\\% - 75\\% above that of the comparison proton model where the simulations match quite closely. As noted in the introduction, the same analysis applies to a variety of cases - different interactions and different identities of primary particle. Given that the same cross section input is not unique to the low scale gravity inspiration used here, this result is of quite general use. Whether neutrinos follow the standard model extrapolations \\cite{q+r,Frich96}, are enhanced ``modestly'' by 3-5 orders of magnitude or ``extravagantly'' by more than 5 orders of magnitude, the search for UHE neutrino induced events at present and new facilities \\cite{auger,icecube} will be an exciting one. {\\bf Note Added}: As we were completing this paper, a closely related work appeared \\cite{neu}. The cross sections considered in their work are well below 10 mb, while we concentrate primarily on cross sections above this value. Where cross sections are roughly the same, results and conclusions qualitatively agree. {\\bf Acknowledgments:} DM thanks Seif Randjbar-Daemi and the High Energy Group at Abdus Salam I.C.T.P. and Jack Gunion and the High Energy Group at the Department of Physics, U.C. Davis for sabbatical leave hospitality while this work was being completed. AJ and PJ thank S. J. Sciutto for help in using the AIRES air shower simulator. We also thank Jaime Alvarez for useful discussions. This work was supported in part by U.S. DOE Grant number DE-FG03-98ER41079, the {\\it Kansas Institute for Theoretical and Computational Science} and DST (India) grant No. DST/PHY/19990184." }, "0011/astro-ph0011249_arXiv.txt": { "abstract": "The detection of Gamma Ray Burst GRB990705 on 1999, July 5.66765 UT, pointing to the Large Magellanic Clouds, suggested the search for a possible neutrino counterpart, both in coincidence with and slightly before (or after) the photon burst. We exploited such a possibility by means of the LVD neutrino telescope (National Gran Sasso Laboratory, Italy), which has the capability to study low-energy cosmic neutrinos. No evidence for any neutrino signal, over a wide range of time durations, has been found, at the occurrence of GRB990705. Due to the lack of information about both the source distance and its emission spectrum, the results of the search are expressed in terms of upper limits, at the Earth, to the $\\bar{\\nu}_\\mathrm{e}$ flux $\\cdot$ cross-section, integrated over different time durations, $\\int \\int \\Phi_{\\bar \\nu_\\mathrm{e}}\\sigma dE dt$. Moreover, assuming thermal $\\bar\\nu_\\mathrm{e}$ spectra at the source, upper limits to the $\\bar\\nu_\\mathrm{e}$ flux, integrated over time duration, for different spectral temperatures, are obtained. Based on these limits and on the expectations for $\\nu$ emission from collapsing astrophysical objects, the occurrence of a gravitational stellar collapse can be excluded up to a distance $r \\approx 50$ kpc, in the case of time coincidence with GRB990705, and $r \\approx 20$ kpc, for the 24 hours preceding it. ", "introduction": "Gamma Ray Burst GRB990705 was detected on 1999, July 5.66765 UT, by the BeppoSAX Gamma-Ray Burst Monitor, and localized by the BeppoSAX Wide Field Camera (Celidonio et al., 1999). It was promptly noted (Djorgovski et al., 1999) that its position, in projection, corresponded to the outskirts of the Large Magellanic Cloud (LMC), and it was suggested that, if the burst was indeed located in the LMC or its halo, a search for a neutrino signal, coincident with, or just prior to the GRB, would be quite interesting. At the time of GRB990705, the LVD neutrino observatory, located in the Gran Sasso underground Laboratory, Italy, was regularly taking data, with active scintillator mass $M=573$ tons. The main purpose of the telescope is the search for neutrinos from gravitational stellar collapses in the Galaxy. On July 19$^{\\mathrm{th}}$ 1999, the result of a preliminary analysis of the LVD data recorded during 48 hours around the time of GRB990705 was reported (Fulgione, 1999), and the absence of a neutrino signal, that would be expected from a gravitational stellar collapse in our Galaxy, was established (no additional results from other neutrino observatories were reported). The search for low-energy neutrinos possibly associated to GRBs is indeed of interest, especially in view of the recent observational evidence linking (some) GRBs and supernovae (see, e.g., Galama et al., 1998, Bloom et al., 1999, Reichart, 1999). Many recent widely discussed models of the sources of GRBs involve the core collapse of massive stars (see, e.g., Woosley, 1993, Paczynski, 1998, Mac Fayden \\& Woosley, 1999, Khokhlov et al., 1999, Wheeler et al., 1999): in this scenario the neutrino emission could be associated to the cooling phase of the collapsed object, the time separation between the neutrino and gamma signals depending on the time necessary to transfer energy from the central engine, which emits thermal $\\nu$, to the outer region, emitting high energy photons. It is clear that the possibility of detecting neutrinos correlated to GRBs depends on the distance of the associated source: even if it appears established that most of them lie at cosmological distances (Metzger et al., 1997), there is evidence, for at least one of GRBs, to be related to a supernovae event in the local universe (Tinney et al., 1998). In particular, from the study of the afterglow of GRB990705 (Masetti et al., 2000), although an extragalactic origin might be supported, the association with LMC cannot be ruled out. Consequently, a more careful analysis of the LVD data in correspondence of GRB990705 has been performed, to search for weaker neutrino signals, not only in coincidence with, but also preceding\\footnote{ By analogy with SN explosions modelling, where few hours are required for the shock to reach the star envelope and give rise to the sudden increase of luminosity, a similar time gap can be assumed, between neutrinos and high-energy $\\gamma$-rays. } and even shortly following it. The paper is planned as follows: in Sect.2 we briefly describe the LVD detector, and we explain the structure of the data. In Sect.3 we present the results of the analysis: a search for a $\\bar{\\nu}_\\mathrm{e}$ signal coincident in time with GRB990705 has been performed. Moreover, a time interval spanning from 24 hrs preceding the burst up to 10 minutes later, has been scanned, searching for any non-statistical fluctuation of the background. For sake of completeness, a wider interval, since 10 days before to 1 day after the event, has been investigated. We conclude in Sect.4, discussing the results in terms of upper limits to the $\\bar \\nu_\\mathrm{e}$ flux possibly associated to the GRB, under the hypothesis of thermal neutrino energy spectrum at the source, and comparing such limits with the expectations from existing models on $\\nu$ emission from collapsing objects. ", "conclusions": "% The number of expected $\\bar\\nu_\\mathrm{e}$ interactions, $N_{\\mathrm{ev}}$, in a time interval $\\delta t$, due to a pulsed $\\bar\\nu_\\mathrm{e}$ emission, is defined as: $$N_{\\mathrm{ev}} = M \\cdot N_\\mathrm{p} \\cdot \\epsilon \\int\\limits^{\\delta t}_{0} dt \\int\\limits^\\infty_{E_{\\mathrm{min}}} \\frac{d^2\\phi_{\\bar\\nu_\\mathrm{e}}}{dE_{\\bar\\nu_\\mathrm{e}} dt} \\sigma(E_{\\bar\\nu_\\mathrm{e}}) dE_{\\bar\\nu_\\mathrm{e}}$$ where $\\epsilon$ is the detection efficiency, $M$ [ton] is the active scintillator mass, $N_\\mathrm{p}$ is the number of free protons per scintillator ton, $\\sigma(E_{\\bar\\nu_\\mathrm{e}})$ is the neutrino interaction cross section (Vogel, 1984) and $\\frac{d^2 \\phi_{\\bar\\nu_\\mathrm{e}}}{dE_{\\bar\\nu_\\mathrm{e}} dt}$ is the differential neutrino flux at the Earth. In the absence of any information on the source distance and its emission spectrum, we can express the results of the search in terms of upper limits to the flux $\\cdot$ cross-section, integrated over the time duration, at the Earth: $ \\int dt \\int \\frac{d^2\\phi}{dE dt} \\sigma dE$. These limits, calculated at 90\\% c.l., are reported in Table 2, for various burst duration $\\delta t$, and they are expressed in number of interactions per target proton. Any hypothesis on the $\\bar\\nu_\\mathrm{e}$ source spectrum leads to a limit to the time integrated $\\bar\\nu_\\mathrm{e}$ flux at the Earth. Assuming a thermal spectrum, constant during the emission interval $\\delta t$, i.e.: $$\\frac {d\\Phi_{{\\bar\\nu_\\mathrm{e}}}}{dE_{\\bar\\nu_\\mathrm{e}}} \\propto \\frac {(\\frac{E_{\\bar\\nu_\\mathrm{e}}}{T_{\\bar \\nu_\\mathrm{e}}})^2} {1+exp(- \\frac{E_{\\bar\\nu_\\mathrm{e}}}{T_{\\bar \\nu_\\mathrm{e}}})}$$ upper limits to the time integrated $\\bar\\nu_\\mathrm{e}$ flux are obtained, as a function of the neutrinosphere emission temperature $T_{\\bar \\nu_\\mathrm{e}}$ [MeV]. These results are shown in Fig.\\ref{fig:l2}, for burst duration $\\delta t \\leq 10$ s. \\begin{figure} \\mbox{\\epsfig{file=ms10019.f3,height=8cm,width=8cm}} % \\caption{Upper limits ($90 \\%$ c.l.) to the time integrated $\\bar\\nu_\\mathrm{e}$ flux, at the Earth, for thermal $\\bar \\nu_\\mathrm{e}$ spectra and $\\delta t \\leq 10$ s, compared with expectations for different source distances.} \\label{fig:l2} \\vspace{-0.5cm} \\end{figure} Most theoretical models on the $\\bar\\nu_\\mathrm{e}$ emission from gravitational stellar collapses (Burrows, 1992) predict that the neutron star binding energy, $E_\\mathrm{b} \\approx 3 \\cdot 10^{53}$ erg, is emitted in neutrinos of every flavour (energy equipartition) with thermal energy spectra, during a time interval $\\delta t \\approx 10$ s. The corresponding $\\bar\\nu_\\mathrm{e}$ fluxes at the Earth, calculated, under the approximation of isotropical emission and pure Fermi-Dirac spectrum, for two different source distances: $50$ kpc (i.e., corresponding to the LMC\\footnote {One can compare these results with the neutrino flux observed from SN1987A, which was definitely located in the LMC. According to the combined analysis of the events detected by the KamiokandeII and IMB detectors (Jegerlehner et al., 1996), which yields a total emitted energy $E_\\mathrm{b} = 3.4 \\cdot 10^{53}$ erg and a $\\bar\\nu_\\mathrm{e}$ spectral temperature $T_{\\bar \\nu_\\mathrm{e}} = 3.6$ MeV, the resulting $\\bar\\nu_\\mathrm{e}$ flux at the Earth, integrated over time, is $\\Phi(\\bar\\nu_\\mathrm{e})\\cdot \\delta t \\sim 9. \\cdot 10^{9} \\mathrm {cm}^{-2}$ } ) and $20$ kpc (i.e., corresponding to the outskirts of our Galaxy), are reported in Fig.\\ref{fig:l2} and are compared with the results of the burst search. The occurrence of a gravitational stellar collapse, with $\\bar \\nu_\\mathrm{e}$ emitted in the temperature range $T_{\\bar\\nu_\\mathrm{e}} > 2$ MeV, can then be excluded within a region of radius $r\\approx50$ kpc, in the case of time coincidence with the GRB990705 event, and $r\\approx20$ kpc, for the 24 hours preceding the GRB time\\footnote{ A possible effect of neutrino mixing on the signal from a gravitational stellar collapse would result in the merging of the energy spectra of neutrinos of different flavours. Because we are dealing with electron antineutrinos, which are characterized by a spectral temperature lower then the one of $\\bar\\nu_{\\mu}$ and $\\bar\\nu_{\\tau}$, neutrino oscillation effects would lead to a hardening of the $\\bar\\nu_\\mathrm{e}$ spectrum and, after all, to an increase of the $\\bar\\nu_\\mathrm{e}$ detection probability. Therefore, excluding oscillations into sterile neutrinos, the limits obtained in this work would remain valid even in the case of neutrino mixing.}." }, "0011/astro-ph0011280_arXiv.txt": { "abstract": "We present a list of 34 neglected entries from star cluster catalogues located at relatively high galactic latitudes ($|b| >$ 15$^{\\circ}$) which appear to be candidate late stages of star cluster dynamical evolution. Although underpopulated with respect to usual open clusters, they still present a high number density contrast as compared to the galactic field. This was verified by means of (i) predicted model counts from different galactic subsystems in the same direction, and (ii) Guide Star Catalog equal solid angle counts for the object and surrounding fields. This suggests that the objects are physical systems, possibly star clusters in the process of disruption or their fossil remains. The sample will be useful for followup studies in view of verifying their physical nature. ", "introduction": "Star Clusters are known to dynamically evolve and stellar depletion effects eventually lead to the cluster dissolution. Fundamental questions are: (i) where are the clusters in process of dissolution? (ii) if fossils are left, can any remnant be detected? The present study aims at showing that several candidates for these effects occur in star cluster catalogues themselves. Several poorly populated objects at relatively high galactic latitudes (from the open cluster perspective $|b| >$ 15$^{\\circ}$) are included in open cluster catalogues (Alter et al. 1970, Lyng\\aa \\,1987). Other objects were reported as clusters in early studies (e.g. New General Catalogue and Index Catalogue), or in modern ones like the ESO (B) Atlas Survey Catalogue (Lauberts 1982). Depletion of Main Sequence (MS) stars has been detected or evidence of it has been found in luminosity functions and Colour-Magnitude Diagrams (CMDs) of some Palomar or Palomar-like globular clusters such as E3 (McClure et al 1985), ESO452-SC11 (Bica et al. 1999) and NGC6717 (Palomar 9) (Ortolani et al. 1999). Such evolved dynamical stages of low mass globular clusters are still associated to relatively well populated star clusters, but one may wonder what subsequent stages would look like, probably an underpopulated fossil core containing some double and multiple stars. Among open clusters low MS depletion has been found e.g. in the intermediate age ($\\approx$ 3 Gyr) cluster NGC3680 (Anthony-Twarog et al. 1991). The dissolution of open clusters has been studied by Wielen (1971). Updated data suggest that most open clusters dissolve in about 100 Myr and this will probably not change much as one includes fainter clusters (Ahumada et al. 2000). Intermediate age open clusters are certainly survivors of initially massive clusters (Friel 1995), and their updated age histogram containing more than 100 entries (Dutra \\& Bica 2000) suggests a dissolution timescale of about 1 Gyr. Also, N-body simulations of star clusters in an external potential have shown typical dissolution times in the range 500-2500 Myr (Terlevich 1987, McMillan \\& Hut 1994, de la Fuente Marcos 1997, Portegies Zwart et al. 2000). Mass segregation is expected to occur in cluster cores during one relaxation time, according to N-body simulations (Terlevich 1987, Portegies Zwart et al. 2000). An important effect of mass segregation is the depletion of low mass MS stars by means of evaporation due to the tidal field of the Galaxy and encounters with binary stars. This would imply that clusters which are close to disruption have a core rich in compact and giant stars (Takahashi \\& Portegies Zwart 2000). Recently, evidence of an open cluster remnant has been discussed by Bassino et al. (2000). They studied the relatively high latitude concentration of stars M73 (NGC6994) and derived an age of 2-3 Gyr from CMDs. They found a significant number density contrast with respect to the galactic field CMD predicted by count models in the area. Carraro (2000) does not favour the object as an open cluster or as a remnant. At any rate, if M73 is a physical system the open cluster classification is certainly not adequate. Let us then suggest the acronym POCR - Possible Open Cluster Remnant. As an ongoing study of this neglected class of interesting objects and in view of future CMDs to determine parameters such as reddening, age and distance, we present a list of candidates. We discuss their possible physical nature by checking whether they present a significant number density contrast with respect to their fields. In Section 2 we present the sample. In Section 3 we analyze the significance of the excesses of stars by means of (i) equal solid angle counts in the object and field areas, and (ii) galactic model counts. In Section 4 we discuss the results. Finally, in Section 5 we present the concluding remarks. ", "conclusions": "We presented a list of 34 neglected entries from star cluster catalogues located at relatively high galactic latitudes ($|b| >$ 15$^{\\circ}$) which may be late stages of cluster dynamical evolution. Although underpopulated with respect to usual open clusters, we showed that they still present a high number density contrast with respect to the galactic field, as verified by means of (i) predicted model counts from different galactic subsystems in the same direction, and (ii) Guide Star Catalog equal solid angle counts for the object and surrounding fields. The sample will be useful for followup studies, aimed at verifying their physical nature. Photometry and spectroscopy are required to determine fundamental parameters such as reddening, distance, age, radial velocities, membership and chemical abundances. Some of these objects might be clusters in the process of disruption or their fossil remnants. The dynamical state of the physical objects in the present sample may be inferred from comparisons of the cluster remnant and surrounding field luminosity functions, searching for depletion effects. An important population of possible open cluster remnants is likely to exist. They may survive significant amounts of time as depopulated systems before dissolution. Simple arguments based on available numerical models and catalogued open clusters suggest that several hundreds can be expected. Systematic surveys to find new candidates and numerical models to explore in detail the evolved dynamical stages are encouraged for a better understanding of this so far overlooked object class." }, "0011/astro-ph0011555_arXiv.txt": { "abstract": "We present an overview of the upcoming Microwave Anisotropy Probe (MAP) mission, with an emphasis on those aspects of the mission that simplify the data analysis. The method used to make sky maps from the differential temperature data is reviewed and we present some of the noise properties expected from these maps. An overview of the method we plan to use to mine the angular power spectrum from the mega-pixel sky maps closes the paper. ", "introduction": "In 1992 NASA's Cosmic Background Explorer (COBE) satellite made a full sky map of the cosmic microwave background (CMB) temperature with 7$^\\circ$ resolution, uncorrelated pixel noise, minimal systematic errors, and accurate calibration, from which CMB temperature anisotropy was first discovered \\cite{smoot92}, \\cite{bennett92}, \\cite{wright92}, \\cite{bennett96}. The purpose of the MAP mission is to re-map the anisotropy over the full sky with more than 30 times the angular resolution ($\\sim 0.23^{\\circ}$ FWHM) and more than 35 times the sensitivity ($\\sim 20$ $\\mu$K per 0.3$^{\\circ}$ pixel) of COBE, but with the same level of quality control as was possible with COBE. With this data, MAP will measure the physical interactions of the photon-baryon fluid (sound waves) in the early universe and thereby test models of structure formation, the geometry of the universe, and inflation. In the years since COBE, a host of ground-based and balloon-borne experiments have detected and characterized fluctuations at smaller angular scales, most recently the experiments TOCO \\cite{toco}, BOOMERanG \\cite{boomldb} and MAXIMA \\cite{maxima}. However, because of their proximity to the Earth and its atmosphere, none of the ground or balloon-based experiments enjoy the extent of systematic error rejection or calibration accuracy that was possible with COBE. Moreover, many of these experiments have significantly correlated noise that places severe demands on the data analysis. The COBE data still serves as a benchmark for the field and many aspects of the COBE mission have influenced the design of the MAP mission. The need to minimize the level of systematic errors has been the major driver of the MAP design and has led to the following high level design features: \\begin{itemize} \\item {\\bf A highly symmetric differential design}: MAP is a differential experiment based on pseudo-correlation microwave radiometers that employ phase-matched HEMT amplifiers. The instrument measures temperature differences between two points $\\sim$ 141$^{\\circ}$ apart on the sky. By measuring temperature differences, rather than absolute temperatures, many spurious signals will be common-mode and thus cancel upon differencing. Also, by employing a pseudo-correlation design with a fast chopping frequency between the two sky inputs, 1/f noise that arises from the HEMTs can be chopped out. The resulting power spectrum of the radiometer noise is very nearly white (see \\S 3). \\item {\\bf Multi-frequency}: There are five frequency bands from 22-90 GHz that will allow emission from the Galaxy and other non-cosmological sources to be modeled and removed based on their frequency dependence. In the lowest frequency bands, MAP will probe the high frequency tail of radio emission from our Galaxy and provide valuable data on the enigmatic microwave foreground emission that correlates with thermal dust emission, but has a much different spectrum. See \\cite{kogut} for a recent summary of evidence for this foreground. \\item {\\bf Stability}: MAP will observe from a Lissajous orbit about the L2 Lagrange point 1.5 million km from Earth. The L2 point offers an exceptionally stable environment and an unobstructed view of deep space, with the Sun, Earth, and Moon always in shadow behind MAP's Sun shield. MAP's large distance from Earth protects it from near-Earth emission and other disturbances. While observing at L2, MAP's Sun shield and solar panels maintain a fixed angle with respect to the Sun to provide exceptional thermal and power stability. \\item {\\bf Low beam sidelobe levels}: The MAP optical system was designed with the foremost goal of providing adequate angular resolution along the line of sight while at the same time rejecting stray light from other directions. For example, the largest instantaneous signal due to radiation from the galactic plane spilling into a sidelobe is expected to be less than 2 $\\mu$K at 90 GHz. \\end{itemize} An overview of the MAP satellite is shown in Fig.~\\ref{map_view}. The major visible features of MAP include the back-to-back telescope optics with 1.4 $\\times$ 1.6 m primary mirrors and 1 m secondary mirrors, the passive thermal radiators which cool the HEMT amplifiers to $<$100 K, the hexagonal structure housing the spacecraft service modules, and the large solar panel array/Sun shield which keeps the instrument in full shade. MAP weighs a total of 830 kg and stands $\\sim$4 m tall. It will be launched in 2001 aboard a Delta 7425-10 expendable launch vehicle from the NASA Kennedy Space Center Eastern Test Range. \\begin{figure} \\centering \\includegraphics*[width=0.8\\hsize,keepaspectratio]{hinshawf1.eps} \\caption{\\small An overview of the MAP satellite.} \\label{map_view} \\end{figure} ", "conclusions": "" }, "0011/astro-ph0011139_arXiv.txt": { "abstract": "We investigate current problems in obtaining reliable ages for old stellar systems based on stellar population synthesis modelling of their integrated spectra. In particular, we address the large ages derived for the globular cluster 47~Tuc, which is at odds with its Color-Magnitude-Diagram (CMD) age. Using a new age indicator, H$\\gamma_{\\sigma<130}$, which is particularly effective at breaking the degeneracy between age and metallicity, we confirm the discrepancy between the spectroscopic age and the CMD age of 47~Tuc, in that the spectroscopic age is much older. Nebular emission appears unlikely to be a source for weakening the observed Balmer lines. We then explore a number of key parameters affecting the temperature of Turn-Off stars, which are the main contributors to the Balmer lines for old metal-rich stellar populations. We find that $\\alpha$-enhanced isochrones with atomic diffusion included not only provides a good fit to the CMD of 47~Tuc, but also leads to a spectroscopic age in better agreement with the CMD age. ", "introduction": "An estimate of the mean luminosity-weighted stellar age of an early-type galaxy represents a major step in unveiling its true star formation history. However, to derive reliable information about stellar ages from the integrated light of unresolved galaxies one must deal with the age-metallicity degeneracy problem, which affects not only integrated colours but also absorption line-strengths (Worthey 1994). Recently, new age-dating techniques based on the Balmer lines (e.g., Jones \\& Worthey 1995; Vazdekis \\& Arimoto 1999, thereafter VA99) have shown great promise in untangling the age-metallicity degeneracy. These techniques should be tested and calibrated on the metal-rich Galactic globular clusters (GCs) for which, unlike elliptical galaxies, independent age estimates are possible (Gibson et al. 1999, thereafter G99) by means of the Color-Magnitude-Diagram (CMD) of their resolved stellar population. The application of the new age-dating techniques to very high signal-to-noise ratio (S/N) spectra of metal-rich Galactic GCs has revealed two major concerns: {\\it i)} the obtained ages are unreasonably large ($>$20~Gyr) (Jones 1999; Cohen, Blakeslee \\& Ryzhov 1998; G99; VA99) and {\\it ii)} a severe disagreement is found between the spectroscopic and CMD ages of 47~Tuc (G99). The CMD-derived ages may be sensitive to the dating method employed (e.g., Alonso et al. 1997) and to the input physics of the theoretical isochrones used as a reference (e.g., Salaris \\& Weiss 1998, thereafter SW98). However, a variety of recent CMD-based age determinations for 47~Tuc have consistently found its age to lie within 9-12.5 Gyr (SW98; Gratton et al.~1997; Carretta et al.~2000; Liu \\& Chaboyer~2000). These CMD-derived ages are substantially younger than the $>20$~Gyr spectroscopically derived age obtained by G99 using the age indicator of Jones \\& Worthey (1995), and Worthey (1994) stellar population models. This discrepancy shows clearly that current stellar population synthesis models used for interpreting the integrated light of stellar systems may have severe zero points problems. In \\S~2 we improve the spectroscopic age-dating technique which confirms a large spectroscopic age for 47~Tuc. In \\S~3 we discuss the possible origin of the problem, explore several theoretical parameters and suggest a possible solution. Finally, in \\S~4 we present our conclusions. ", "conclusions": "We have discussed the origin of the discrepancy between the spectroscopic (based on the effective temperature of TO stars) and CMD (based on the luminosity of TO stars and an assumed distance scale) age estimate for 47~Tuc as raised by G99. For this purpose we have defined a new age indicator, H$\\gamma_{\\sigma<130}$, particularly suitable for studying GCs and low velocity dispersion galaxies, which shows a superb power to break the age-metallicity degeneracy. H$\\gamma_{\\sigma<130}$ confirms the age discrepancy found by G99 for 47~Tuc. Emission fill-in of the Balmer lines appears to be an unlikely source of the weak H$\\gamma$ in 47~Tuc, since the ages derived from different Balmer lines give discordant results if the hypothetical emission fill-in is corrected for, and since a composite of four other metal-rich Galactic globular clusters shows the same weak H$\\gamma$ phenomenon. Thus the fact that other metal-rich GCs show very similar low Balmer values in comparison to the model predictions suggests a problem in the zero point of current stellar population models. It is worth noting that this zero point problem of the models with respect to the metal-rich GCs also works out for old elliptical galaxies. We therefore analyzed the possible causes of the problem by studying a number of input parameters of the evolutionary computations, and comparing the observed value of H$\\gamma_{\\sigma<130}$ with that derived from the synthesized integrated spectra. Neither the initial He content nor the HB have significant effects on the Balmer indices synthesized for 47~Tuc. However the inclusion of $\\alpha$-elements enhancement and atomic diffusion in the evolutionary models provide spectroscopic ages which are much closer to the CMD derived ages. This occurrence constitutes a possible solution to the age-discrepancy between CMD and integrated spectrum ages of old metal-rich stellar populations. It is important to study if the age discrepancy is present in metal poor GCs. For this purpose we need to expand the current stellar spectral libraries which feed the stellar population models (see V99), to extend their predictions to lower metallicities." }, "0011/astro-ph0011413_arXiv.txt": { "abstract": "We present HST images at 622\\,nm and 300\\,nm of the jet in 3C\\,273 and determine the run of the optical spectral index at 0\\farcs2 along the jet. We find no evidence for localized acceleration or loss sites, and support for a little-changing spectral shape throughout the jet. We consider this further evidence in favour of a distributed acceleration process. ", "introduction": " ", "conclusions": "" }, "0011/astro-ph0011375_arXiv.txt": { "abstract": "We present an update of results from the search for microlensing towards the Large Magellanic Cloud (\\lmc) by \\eros\\ (Exp\\'erience de Recherche d'Objets Sombres). We have now monitored 25 million stars over three years. Because of the small number of observed microlensing candidates (four), our results are best presented as upper limits on the amount of dark compact objects in the halo of our Galaxy. We discuss critically the candidates and the possible location of the lenses, halo or \\lmc . We compare our results to those of the {\\sc macho} group. Finally, we combine these new results with those from our search towards the Small Magellanic Cloud as well as earlier ones from the {\\sc eros1} phase of our survey. The combined data is sensitive to compact objects in the broad mass range $10^{-7} - 10 \\, {\\rm M}_{\\odot}$. The derived upper limit on the abundance of stellar mass {\\sc macho}s rules out such objects as the dominant component of the Galactic halo if their mass is smaller than $2 {\\rm M}_{\\odot}$. ", "introduction": "The search for gravitational microlensing in our Galaxy has been going on for a decade, following the proposal to use this effect as a probe of the dark matter content of the Galactic halo \\cite{pac86}. The first microlensing candidates were reported in 1993, towards the \\lmc\\ \\cite{aub93,alc93} and the Galactic Centre \\cite{uda93} by the \\eros , \\macho\\ and \\ogle\\ collaborations. Because they observed no microlensing candidate with a duration shorter than 10~days, the \\erou\\ and \\macho\\ groups were able to exclude the possibility that more than 10\\% of the Galactic dark matter resides in planet-sized objects \\cite{aub95,alc96,ren97,ren98,alc98}. However a few events were detected with longer time\\-scales. In their two-year analysis \\cite{alc97a}, the \\macho\\ group estimated, from 6-8 candidate events towards the \\lmc , an optical depth of order half that required to account for the dynamical mass of a ``standard'' spherical dark halo\\footnote{ $4 \\times 10^{11}\\:{\\rm M}_\\odot$ within 50~kpc}; the typical Einstein radius crossing time of the events, $t_E$, implied an average mass of about 0.5~M$_\\odot$ for halo lenses \\cite{alc97a}. Based on two candidates, \\erou\\ set an upper limit on the halo mass fraction in objects of similar masses \\cite{ans96,ren97}, that is below that required to explain the rotation curve of our Galaxy\\footnote{ Assuming the original two \\erou\\ candidates are microlensing events, they would correspond to an optical depth six times lower than that expected from a halo fully comprised of \\macho s.}. The second phase of the \\eros\\ programme was started in 1996, with a ten-fold increase over \\erou\\ in the number of monitored stars in the Magellanic Clouds. The analysis of the first two years of data towards the Small Magellanic Cloud (\\smc ) allowed the observation of one microlensing event \\cite{pal98} also detected by \\cite{alc97b}. This single event, out of 5.3 million monitored stars, allowed \\erod\\ to further constrain the halo composition, excluding in particular that more than 50~\\% of the standard dark halo is made up of $0.01 - 0.5 \\:{\\rm M}_\\odot$ objects \\cite{afo99}. In contrast, an optical detection of a halo white dwarf population was reported \\cite{iba99}, compatible with a galactic halo full of white dwarfs. Very recently, the \\macho\\ group presented an analysis of 5.7 year light curves of 10.7~million stars in the \\lmc\\ \\cite{alc00} with an improved determination of their detection efficiency and a better rejection of background supernova explosions behind the \\lmc . They now favour a galactic halo \\macho\\ component of 20\\% in the form of 0.4~M$_\\odot$ objects. Within a few days, the detection of a halo white dwarf population at the level of a 10\\% component was also reported by \\cite{iba00}. Simultaneously, the \\erod\\ group presented its results from a two-year survey of 17.5 million stars in the \\lmc\\ \\cite{las00a}. One \\erou\\ microlensing candidate, {\\sc eros1-lmc-2}, was seen to vary again, 8 years after its first brightening, and was thus eliminated from the list of microlensing candidates. Two new candidates were identified ({\\sc eros2-lmc-3} and~4). Because this is much lower than expected if \\macho s are a substantial component of the galactic halo, and because these two new candidates do not show excellent agreement with simple microlensing light curves, \\eros\\ chose to combine these results with those from previous \\eros\\ analyses, and to quote an upper limit on the fraction of the galactic halo in the form of \\macho s. In this article, we describe an update on the \\erod\\ \\lmc\\ data, an analysis of the three-year light curves from 25.5 million stars. While the sensitivity is improved, the main conclusions are unchanged compared to \\cite{las00a}. One of the two-year candidates was seen to vary in the third season and was thus rejected. Three new candidates have been detected. We combine these \\erod\\ \\lmc\\ results with those of previous independent \\eros\\ analyses, and derive the strongest limit obtained thus far on the amount of stellar mass objects in the Galactic halo. ", "conclusions": "After nine years of monitoring the Magellanic Clouds, \\eros\\ has a meager crop of five microlensing candidates towards the \\lmc\\ and one towards the \\smc , whereas about 30 events are expected towards the \\lmc\\ for a spherical halo fully comprised of $0.4 \\:{\\rm M}_\\odot$ objects. Moreover, some of the candidates cannot be considered excellent. These candidates were obtained from four different data sets analysed by independent, cross-validated programs. So, the small number of observed events is unlikely to be due to bad (and overestimated) detection efficiencies. This allows us to put strong constraints on the fraction of the halo made of objects in the range [$10^{-7}\\:{\\rm M}_\\odot$, $10\\:{\\rm M}_{\\odot}$], excluding in particular at the 95\\% C.L. that more than 40\\% of the standard halo be made of objects with up to $1 \\:{\\rm M}_\\odot$. The preferred value quoted in \\cite{alc00}, $f = 0.2$ and $0.4\\:{\\rm M}_\\odot$, is consistent with our limit as can be seen in Fig.~\\ref{excl}. (The upper part -~about 25\\%~- of the domain allowed by \\cite{alc00} is excluded by the limit we report here.) There are several differences which should be kept in mind while comparing the two experiments. First, \\eros\\ uses less crowded fields than \\macho\\ with the result that blending is relatively unimportant for \\eros . (Were \\eros\\ results to be corrected for blending, the detection efficiency would increase slightly and the reported limit would be stronger.) Second, \\eros\\ covers a larger solid angle (64~deg$^2$ in the \\lmc\\ and 10~deg$^2$ in the \\smc ) than \\macho , which monitors primarily 15~deg$^2$ in the central part of the \\lmc . The \\eros\\ rate should thus be less contaminated by self-lensing, i.e. microlensing of \\lmc\\ stars by dimmer \\lmc\\ objects, which should be more common in the central regions. The importance of self-lensing was first stressed in \\cite{wu94,sa94}. Third, the \\macho\\ data have a more frequent time sampling. The results from \\eros\\ and \\macho\\ are apparently consistent, but the way they are interpreted is different. \\macho\\ reports a signal and considers the contamination of its sample as low or null. \\erod\\ quotes an upper limit and does not claim its sample to be background-free. The position of the lenses along the line of sight, halo or Magellanic Clouds, is also an issue. \\macho\\ has compared the spatial distribution of its candidates across the face of the \\lmc\\ and observes a better agreement with the halo hypothesis than with a specific model of the \\lmc . On the other hand, because the \\eros\\ stars are spread over a wider field, the fact that the \\eros\\ sample corresponds to a lower central value of the event rate (about twice lower than that of \\macho ) is compatible with an interpretation where a noteable fraction of the events are due to self-lensing. The small number of \\eros\\ candidates precludes at present any definitive conclusion on that topic. It seems likely that the single most important input to the question of the position of the lenses will come from the comparison of the microlens candidates samples towards the \\smc\\ with those towards the \\lmc . Because the two lines of sight are rather close (about 20~degrees apart), the timescale distributions of microlensing candidates towards the two Clouds should be nearly identical if lenses belong to the galactic halo. Also, the event rates should be comparable, although the ratio is more halo model dependent. At present, \\eros\\ has analysed two seasons of \\smc\\ data \\cite{afo99} and \\macho\\ has not yet presented its detection efficiency towards the \\smc . From the published \\eros\\ efficiencies, and assuming that the \\macho\\ efficiencies towards the \\smc\\ are similar to those towards the \\lmc , it can be expected that the completed experiments will have gathered between five and ten microlenses towards the \\smc . This should allow a significant comparison of the timescales (see also the discussion in \\cite{gra00}). Finally, let us mention that, given the scarcity of our candidates and the possibility that some observed microlenses actually lie in the Magellanic Clouds, \\eros\\ is not willing at present to quote a non zero {\\it lower} limit on the fraction of the Galactic halo comprised of dark compact objects with masses up to a few solar masses. \\bigskip {\\bf Acknowledgements} We are grateful to D. Lacroix and the staff at the Observatoire de Haute Provence and to A. Baranne for their help with the MARLY telescope. The support by the technical staff at ESO, La Silla, is essential to our project. We thank J.F. Lecointe for assistance with the online computing. We thank all non-\\eros\\ members who participated in data taking." }, "0011/astro-ph0011143_arXiv.txt": { "abstract": "\\emph{Eddington} is a space mission for extrasolar planet finding and for asteroseismic observations. It has been selected by ESA as an F2/F3 reserve mission with a potential implementation in 2008-13. Here we describe \\emph{Eddington}'s capabilities to detect extrasolar planets, with an emphasis on the detection of habitable planets. Simulations covering the instrumental capabilities of \\emph{Eddington} and the stellar distributions in potential target fields lead to predictions of about 10,000 planets of all sizes and temperatures, and a few tens of terrestrial planets that are potentially habitable. Implications of \\emph{Eddington} for future larger scale missions are briefly discussed. ", "introduction": "mission} The \\emph{Eddington} mission is a space telescope designed for two primary goals: asteroseismic studies and extrasolar planet finding. Both goals will be achieved through the acquisition of high-precision wide-field photometry, with a temporal stability that is possible only from space. The basic design is an f/3 triple-reflecting telescope with a 1.2m diameter aperture, 0.6m$^2$ effective area, and a 3$\\deg$ diameter field of view imaged by a 20-CCD mosaic camera. The telescope will be launched into an orbit around the L2 Earth-Sun libration point, from which the entire sky is accessible for observations. The first 2 years of the mission will be dedicated primarily to asteroseismic studies, with pointings lasting up to two months to a variety of targets. The mission's second part will be dedicated to planet finding by the transit method, where the telescope will survey a single stellar field for 3 years. In total, about 500,000 stars will be surveyed by \\emph{Eddington} as potential hosts for planetary systems. In October 2000, \\emph{Eddington} was selected by ESA as a reserve mission for the F2/F3 launch window (2009-2013). More details about the mission can be found in the \\emph{Eddington} Assessment study (Favata et al., 2000) and in the \\emph{Eddington} web-site (http://astro.esa.int/SA-general/Projects/Eddington/). Here we will give a short overview about its planet detection capabilities. ", "conclusions": "" }, "0011/astro-ph0011469_arXiv.txt": { "abstract": "We report the first results from the BOOMERanG experiment, which mapped at 90, 150, 240 and 410 GHz a wide ($3\\%$) region of the microwave sky with minimal local contamination. From the data of the best 150 GHz detector we find evidence for a well defined peak in the power spectrum of temperature fluctuations of the Cosmic Microwave Background, localized at $\\ell = 197 \\pm 6$, with an amplitude of $(68 \\pm 8) \\mu K_{CMB}$. The location, width and amplitude of the peak is suggestive of acoustic oscillations in the primeval plasma. In the framework of inflationary adiabatic cosmological models the measured spectrum allows a Bayesian estimate of the curvature of the Universe and of other cosmological parameters. With reasonable priors we find $\\Omega = (1.07 \\pm 0.06)$ and $n_s = (1.00 \\pm 0.08)$ (68$\\%$C.L.) in excellent agreement with the expectations from the simplest inflationary theories. We also discuss the limits on the density of baryons, of cold dark matter and on the cosmological constant. ", "introduction": " ", "conclusions": "" }, "0011/astro-ph0011233_arXiv.txt": { "abstract": "Chiral cosmic strings are naturally produced at the end of D-term inflation and they may have interesting cosmological consequences. As was first proved by Carter and Peter, the equations of motion for chiral cosmic strings in Minkowski space are integrable (just as for Nambu-Goto strings). Their solutions are labeled by a function $k(\\sigma - t)$ where $t$ is time and $\\sigma$ is the invariant length along the string, and the constraints on $k$, which determines the charge on the string, are that $0 \\leq k^2 \\leq 1$. We review the origin of this parameter and also discuss some general properties of such strings which can be deduced from the equations of motion. The metric around infinite chiral strings is then constructed in the weak field limit, and studied as a function of $k$. We also consider the angular momentum of circular chiral loops, and extend previous work to consider the evolution and self-intersection properties of a more general family of chiral cosmic string loops for which $k^2(\\sigma-t)$ is not constant. ", "introduction": "In the last few years the scenario of structure formation from cosmic strings has become increasingly tenuous, since its predictions differ significantly from the new high accuracy measurements of the temperature fluctuations in the cosmic microwave background radiation. Most studies of such observational consequences of strings have focused on structureless Nambu-Goto (NG) strings \\dc{James1,CHM,avelino,Levon,CMS} and global strings \\dc{Neil,Ruth}, and in each case the recent predictions are based on numerical simulations of the evolution of the string network postulated to form at the GUT phase transition. One should recall though that there are some unresolved and potentially important uncertainties in the simulations --- it is very difficult, for example, to resolve the very disparate scales which characterize the the network, as well as to deal with gravitational backreaction effect --- and hence a combination of numerical work with analytical modeling \\dc{James1,Levon,CMS} has also been used to make predictions from NG strings. Our focus here is not on NG strings but rather on {\\em chiral cosmic strings}. These strings are a type of current carrying string \\dc{Witten} for which the world-sheet current $j^{i}$ is null; $$ j^{i}j_i = j^2 = 0. $$ (Here $i=(0,1)$ and the 2D world sheet metric $\\gamma_{ij}$ defined below raises and lowers indices.) One motivation for studying such chiral strings comes from the well known SUSY D-term inflation model. In this model, strings are produced at the end of inflation \\dc{Rachel} so that both mechanisms contribute to producing density fluctuations. However, the strings produced are chiral cosmic strings and not NG strings \\dc{DDT1}. Hence in order to make predictions for the $C_l$'s from this `strings plus inflation' model, the evolution and cosmological consequences of chiral cosmic string networks must be understood. (There may exist models in which the strings formed at the end of inflation are NG ones, however this is not true of D-term inflation. In the case of `inflation plus NG strings', predictions may be found in \\dc{CHM2}.) There are a number of differences between the properties of chiral cosmic strings and NG strings. One such regards the evolution of the strings themselves: the null current on chiral strings can, as in the case of other current-carrying strings, lead to the formation of non self-intersecting stable loops called vortons\\footnote{As will be come clearer later, by a vorton we mean a stable loop of arbitrary shape that never self-intersects. This definition is different from that of Martins and Shellard \\dc{MS} who also require that these loops move non-relativistically, suggesting that otherwise the charge on the loops could be `thrown off'. We are not able to comment on such a mechanism, however see \\dc{DDDD} for a discussion of the scattering of zero-modes from chiral strings.}. This is potentially catastrophic as the energy density in the chiral string network could quickly dominate the energy density in the universe if stable vortons are present. It is therefore important to see if vortons are produced, and in section \\dr{s:kvar} we study the self-intersection properties of a family of chiral cosmic string loops. Another difference between NG and chiral strings is that these line-like sources of energy generate different metrics about them (section \\dr{s:metric})\\footnote{I am aware that this comment disagrees with one I made in \\dc{Verb}! I would like to thank Patrick Peter and Tanmay Vachaspati for pointing out an error in my previous determination of the metric.}. One might therefore expect them to produce different perturbations in the matter and radiation through which they pass. Recently a number of steps have been made which allow for a quantitative study of chiral cosmic string dynamics. First, a well defined unique 2D effective action exists for these strings \\dc{MS,CP}. From this action it was shown, with suitable gauge choices, that the equations of motion are integrable in Minkowski space \\dc{CP} (see also \\dc{US,BP} for different presentations of the same result). They are \\be \\frac{\\paa^2 \\vecx}{\\paa t^2} - \\frac{\\paa^2 \\vecx}{\\paa \\sigma^2} = 0 \\qquad \\Longrightarrow \\qquad \\vecx(t,\\si) = \\frac{1}{2}[\\veca(t+\\si) + \\vecb(t-\\si)], \\dle{eqnlecon} \\ee where $t$ is background time, and $\\sigma$ measures the invariant length or energy along the string as in the NG case \\dc{US}. The constraints are \\ba \\spr{\\veca}^2 \\= 1, \\dla{c1} \\\\ \\spr{\\vecb}^2 \\!\\!\\!& \\le &\\!\\!\\! 1, \\dla{c2} \\end{eqnarray} where for instance $\\spr\\veca(q)\\equiv d\\veca(q)/dq.$ If one defines \\be k^2:= \\spr{\\vecb}^2 \\dle{kddef} \\ee so that $k=k(t-\\sigma)$, then it can be shown that $k^2$ determines the conserved charge on the string (see also below). Furthermore, if $k = $ constant $=1$ then this charge vanishes as required, since $\\spr{\\vecb}=1$ is just the Nambu-Goto limit. In reference \\dc{US}, the self-intersection properties of chiral cosmic string loops were also studied in the special case of $k =$ constant. In particular the strings were shown never to self-intersect for $k=0$: this case corresponds to maximal charge on the strings and to vorton solutions. Here that work is extended, though we still consider Minkowski space (with metric $\\eta_{\\mu \\nu} = (+,-,-,-)$) throughout. First, for completeness, we indicate in section \\dr{s:2} how the equations of motion (\\dr{eqnlecon})-(\\dr{c2}) are obtained from the chiral action and how the charge mentioned above is defined. This necessarily follows parts of reference \\dc{US} rather closely, though a small error in that paper is corrected. We also compare the chiral charge with the charges used for more general current carrying strings. In section \\dr{s:angmmt} we summarize some properties of chiral cosmic strings which result from the equations of motion. The metric around infinite chiral strings is then studied as a function of $k$ and we comment in possible consequences it may have for structure formation and CMB anisotropies from chiral cosmic strings. In section \\dr{ss:ang}, the effect of angular momentum on the motion of circular loops is considered by looking at the effective potential introduced in \\dc{CPG}. In section \\dr{s:kvar} we investigate the self-intersection properties of loops with non-constant $k$. Finally conclusions are given in section \\dr{s:conc}. ", "conclusions": "\\dle{s:conc} In this paper we have attempted study and clarify a number of points regarding the evolution and gravitational properties of chiral cosmic strings. As was summarized in section \\dr{s:2}, the crucial difference between the equations of motion for NG and chiral cosmic strings is the constraint on the vector $\\spr{\\vecb}$: for NG strings $\\spr{\\vecb}^2(\\eta) = 1$ $\\forall \\eta$, whereas for chiral strings $\\spr{\\vecb}^2(\\eta) (=k^2(\\eta)) \\leq 1$. Equation (\\dr{Cr}) shows that $k^{2}(\\eta)$ determines the charge on the chiral string. We saw in section \\dr{s:gen} that chiral strings with $k=0$ ($\\forall \\eta$) move along themselves and never self-intersect. If the string forms a loop, the energy of this arbitrary shaped vorton is equipartitioned between tension and angular momentum. The charge on the vortons is given by $C = m L_{{\\rm phys}}$ where $ L_{{\\rm phys}}$ is the constant physical length of the vorton. Infinite straight chiral strings were studied in section \\dr{s:metric}. We saw that the energy momentum tensor contains non-diagonal terms $T^{tz} \\neq 0$. These represent the momentum along the string. Furthermore, $T^{tt} \\neq T^{zz}$ (if $k \\neq 1$) which is reminiscent of the situation which occurs with wiggly NG strings. As a consequence of the form of $T^{\\mu \\nu}$, the weak-field metric around the string was shown to contain a $dt \\, dz$ term which means that photons (and relativistic particles) moving near the string are dragged in the direction of the string. We also observed that there is a $k$-dependent deficit angle as well as a $k$-dependent Newtonian potential. Regarding the evolution of a chiral cosmic string network (which could formed at the end of D-term inflation), it is important to understand whether or not the loops can self-intersect and then decay. If they cannot decay, this would lead to a cosmological catastrophe as they would dominate the energy density of the universe. In section \\dr{ss:ang} we studied the effective potential for the motion of a non self-intersecting circular loop for which $0 \\leq k < 1$. In section \\dr{s:kvar} we considered loops with non-constant $k$: the physical reason for which one might expect $k$ not to be constant is that charge will build up as a result of self-intersections, and also fluctuate during the phase transition which forms the string. Analysis of specific form of $k(\\eta)$ (given via (\\dr{bgen})) showed that self-intersection is possible for these loops. The ensuing numerical analysis showed that the self-intersection probability depends on the form of $k(\\eta)$ and is not uniquely determined by the charge $C$ of the loop. This unfortunately suggests that even if one were able to estimate $C$ for the strings in a chiral cosmic string network, this would not be sufficient to determine the self-intersection properties of the loops. As a further problem it still remains to understand the fate of the daughter loops. A number of interesting questions remain to be studied. Regarding the metric (section \\dr{s:metric}), it would be interesting to go beyond the weak-field approximation and also to study carefully the potential cosmological consequences of the $dt \\, dz$ term \\dc{StVa}. This cross-term is the main difference between the metric for NG and chiral strings. Concerning the evolution of a network of chiral cosmic strings it is clear that if the network is formed with $k(\\eta) = 0$ $\\forall \\eta$ and for all strings, then this leads to a cosmological catastrophe: this is the only case in which the answer for $P_{int}$ is unique and zero! --- the strings cannot self-intersect and are frozen. Similar problems occur if this state is reached at anytime during the evolution of the network. This vorton problem was studied in \\dc{CD} where it was noted that the quantum number $C$ should be larger for chiral strings than for strings with time- or space-like currents. However, work still needs to be done to see if $C$ is maximal or not \\dc{DP}. If it is not maximal (i.e.\\ $k \\neq 0$ $\\forall \\eta$) it still remains to understand the ultimate fate of the daughter loops, and hence that of the network itself." }, "0011/astro-ph0011011_arXiv.txt": { "abstract": " ", "introduction": "The black hole candidate GRS~1915+105 is known for its extremely complex X-ray behaviour, which is not observed in any other X-ray source. Belloni et al. (2000) succesfully described this behaviour as transitions between three basic spectral states: the soft states A and B, and the hard state C. GRS 1915+105 was discovered in 1994 to be the first galactic source to show superluminal motions of radio-emitting ejecta (Mirabel \\& Rodriguez 1994). Besides these big radio flares, the source also shows smaller radio oscillation events with periods around 20-40 minutes (Pooley \\& Fender 1997). ", "conclusions": "" }, "0011/astro-ph0011227_arXiv.txt": { "abstract": "We present a spectrophotometric study based on VLT/FORS\\,I observations of one of the most metal-deficient blue compact dwarf (BCD) galaxies known, \\tol\\ ($Z$$\\sim$$Z$$_{\\odot}$/25). The data show that roughly half of the total luminosity of the BCD originates from a bright and compact starburst region located at the northeastern tip of a faint dwarf galaxy with cometary appearance. The starburst has ignited less than 4 Myr ago and its emission is powered by several thousands O7V stars and $\\sim$ 170 late-type nitrogen Wolf-Rayet stars located within a compact region with $\\la$500 pc in diameter. For the first time in a BCD, a relatively strong [Fe\\,V] $\\lambda$4227 emission line is seen which together with intense He II $\\lambda$4686 emission indicates the presence of a very hard radiation field in \\tol. We argue that this extraordinarily hard radiation originates from both Wolf--Rayet stars and radiative shocks in the starburst region. The structural properties of the low-surface-brightness (LSB) component underlying the starburst have been investigated by means of surface photometry down to 28\\ $B$ \\sbb. We find that, for a surface brightness level fainter than $\\sim 24.5$\\ $B$\\ \\sbb, an exponential fitting law provides an adequate approximation to its radial intensity distribution. The broad-band colors in the outskirts of the LSB component of \\tol\\ are nearly constant and are consistent with an age below one Gyr. This conclusion is supported by the comparison of the observed spectral energy distribution (SED) of the LSB host with theoretical SEDs. ", "introduction": "} \\subsection{Photometric data} Images of \\tol\\ in the broad-band filters Bessell $U$, $B$, $R$ were obtained with the {\\sf Fo}cal {\\sf R}educer and low-dispersion {\\sf S}pectrograph (FORS\\,1; see Moehler et al. 1995) attached to the VLT\\ UT1. The exposures were acquired under photometric conditions on May, 17th 1999 during a 6 night run allocated to guaranteed time observations (GTO) at an airmass ranging between 1.48 and 1.64. The seeing was between 0\\farcs 7 and 0\\farcs 9. FORS was operating in the standard imaging mode yielding a final focal ratio of 3.13 and an instrumental scale of 0\\farcs 2 pix$^{-1}$. Photometric zero-points and color-dependent calibration terms were derived from exposures of the standard-star field Mark\\ A (Landolt 1992) taken each night during the GTO run at an airmass$\\approx$1. Airmass-dependent calibration terms were obtained using standard extinction curves and found to agree to a level better than 10\\% with those derived during the commissioning phase\\,II of FORS. The photometric precision is estimated to be $\\sim 0.1$ mag. Reduction has been accomplished in the standard way using the ESO-MIDAS software package. \\subsection{Spectroscopic data} Spectroscopic data for Tol 1214--277 were taken with FORS\\,1 on May, 12th 1999 at an airmass 1.7 and with a seeing between 0\\farcs7 -- 0\\farcs9 FWHM. A 1\\arcsec\\ $\\times$ 180\\arcsec\\ slit was used in conjunction with a grism GRIS$\\_$300V and a GG375 second-order blocking filter. This yields a spatial resolution along the slit of 0\\farcs2 pixel$^{-1}$, a scale perpendicular to the slit of 3 \\AA\\ pixel$^{-1}$, a spectral coverage of 3600 -- 7500 \\AA, and a spectral resolution of $\\sim$ 10 \\AA\\ (FWHM). The total exposure time of 1650 seconds has allowed to reach a signal-to-noise ratio S/N $\\ga$ 50 in the continuum of the bright central part of the BCD and was broken up into two subexposures, 990 and 660 seconds, to allow for an efficient cosmic-ray removal. The slit was oriented in the position angle P.A. = 39$^{\\circ}$ to enable a simultaneous study of the starburst knot and the faint underlying host along its major axis. A spectrum of a He-Ne-Ar comparison lamp was obtained for wavelength calibration. Since no spectrophotometric standard star was observed, calibration was done using flux-calibrated spectra of \\tol\\ obtained previously with the 2.1m KPNO telescope\\footnote{Kitt Peak National Observatory (KPNO) is operated by the Association of Universities for Research in Astronomy (AURA), Inc. under cooperative agreement with the National Science Foundation.}. These were taken on April 2, 1998 at an airmass 2.0. The total exposure time was 3600 seconds, split into 3 subexposures, 1200 seconds each. A 2\\arcsec\\ $\\times$ 180\\arcsec\\ slit was used along with grating No. 9 and a GG375 second-order blocking filter. This yields a spatial resolution along the slit of 0\\farcs69 pixel$^{-1}$, a scale perpendicular to the slit of 2.7 \\AA\\ pixel$^{-1}$, a spectral range 3600 -- 7500 \\AA, and a spectral resolution of $\\sim$ 7 \\AA\\ (FWHM). The spectrophotometric standard star Feige 34 was observed for flux calibration. Data reduction of the 2.1m telescope and VLT spectroscopic observations was carried out using the IRAF software package. This included bias subtraction, cosmic-ray removal, flat-field correction, wavelength calibration and night-sky background subtraction. ", "conclusions": "The main conclusions drawn from our imaging and spectroscopic analysis of deep VLT data of the extremely metal-deficient ($Z$$\\sim$$Z$$_{\\odot}$/25) and nearby ($D$ = 103.9 Mpc) BCD \\tol\\ may be summarized as follows: 1. \\tol\\ undergoes a vigorous burst of star formation having ignited less than 4 Myr ago. The starburst takes place within a bright ($M_B\\!\\sim\\!-16$ mag) compact ($\\la$ 500 pc in diameter) region, giving rise to extended and abundant ionized gas emission with an H$\\beta$ equivalent width of $\\sim$ 320 \\AA. The starburst is powered by several thousands of O7V stars and 170 late-type nitrogen Wolf--Rayet stars. 2. In this very metal-deficient BCD we discover for the first time the high ionization line [Fe V] $\\lambda$4227. Moreover, we detect extraordinarily strong He II $\\lambda$4686 emission with an intensity as high as 5\\% of that of the H$\\beta$ emission line. This implies the presence of a very hard radiation field in \\tol. The intensity ratio $I$([Fe V] $\\lambda$4227) / $I$(He II $\\lambda$4686) in \\tol\\ compares well with that in another extremely metal-poor BCD with [Fe V] $\\lambda$4227 emission, SBS 0335--052, being in both cases larger by more than one order of magnitude than the ratio observed in high-excitation planetary nebulae. While the relative number of WR stars of $N$(WR) / $N$(O + WR) = 0.023 in \\tol\\ is compatible with theoretical predictions, the intensity of the He II $\\lambda$4686 emission line exceeds several times the predictions of standard H II photoionization models, even when the hard radiation component of Wolf-Rayet stars is taken into account. Therefore, we argue that the hard ionizing radiation field in \\tol\\ is produced from the combined effect of massive stars and SN-driven shocks. 3. Star-forming activity in \\tol\\ is confined to the northeastern tip of a cometary dwarf galaxy with an absolute $B$ magnitude $\\ga$ --16 mag and an isophotal size of 7.6$\\times$4.8 kpc at 28 $B$ \\sbb. An exponential fitting law provides a reasonable approximation to the intensity distribution of the stellar LSB host in its outskirts, for $\\mu_B\\ga 24.5$ \\sbb. It fails, however, to properly describe the observed brightness distribution at intermediate and high intensity levels. These are better fitted by an exponential distribution which flattens at small radii, similar to the V-type profiles described by Binggeli \\& Cameron (1991). 4. The radially averaged ($U-B$) and ($B-R$) colors of the LSB host of \\tol\\ are consistent with those for either an instantaneous burst with log ($t$/yr) $\\la$ 8.1, or a continuous star formation between log ($t$/yr) $\\la$ 8.7 and log ($t$/yr) $=$ 7.3. We however cannot definitely exclude the presence of a small fraction of old (age $>$ 1 Gyr) stars due to their intrinsic faintness." }, "0011/astro-ph0011157_arXiv.txt": { "abstract": "To determine the equation of state of the universe, we propose to use a new independent variable $R\\equiv (H_0/c)(d_L(z)/(1+z))$, where $H_0$ and $d_L(z)$ are the present Hubble parameter and the luminosity distance, respectively. For the flat universe suggested from the observation of the anisotropy of cosmic microwave background, the density and the pressure are expressed as $\\rho/\\rho_0=4(df/dR)^2/f^6$ and $p/\\rho_0=-4/3(d^2f/dR^2)/f^5$ where $\\rho_0$ is the present density and $f(R)=1/\\sqrt{1+z(R)}$. In $(R, f)$ plane the sign as well as the strength of the pressure is in proportion to the curvature of the curve $f(R)$. We propose to adopt a Pade-like expression of $f(R)=1/\\sqrt{u}$ with $u\\equiv 1+\\sum\\limits_{n=1}^{N}u_nR^n$. For flat $\\Lambda$ model the expansion up to $N=7$ has at most an error $< 0.2\\%$ for $z < 1.7$ and any value of $\\Lambda$. We also propose a general method to determine the equation of state of the universe which has $N-1$ free parameters. If the number of parameters are smaller than $N-1$, there is a consistency check of the equation of state so that we may confirm or refute each model. ", "introduction": "Recent measurements of the luminosity distance $d_L(z)$ using Type Ia supernovae (\\cite{schmidt98,ries98,perl99}) suggest that accurate $d_L(z)$ may be obtained in the near future. Especially SNAP (\\cite{snap}) will give us the luminosity distance of $\\sim$2000 Type Ia supernovae with an accuracy of a few \\% up to $z\\sim1.7$ every year. On the other hand from the observation of the first Doppler peak of the anisotropy of CMB, it is now suggested that the universe is flat (\\cite{ber00,lange00}), which may be proved in future by MAP and Planck. If the flat universe is the case, the density $\\rho(z)$ and the pressure $p(z)$ can be determined only from $d_L(z)$ in principle (\\cite{naka99}) so that the equation of the state of the universe is uniquely determined. If not, the determination of the present curvature of the universe and the determination of $\\rho(z)$ and $p(z)$ will be coupled in general (\\cite{naka99}). Now let us assume that the universe is flat. Even in this case at least two problems exist: 1) How to express the continuous function $d_L(z)$ which is accurate enough from $z=0$ to $z\\sim 1.7$ using several free parameters; 2) From $d_L(z)$ how to obtain accurate $\\rho(z)$ and $p(z)$, that is, the equation of state of the universe(\\cite{star98,huter99,naka99,saini99,chiba00}). In this paper, we propose to use a new independent variable $R\\equiv (H_0/c)(d_L(z)/(1+z))$ instead of $z$, where $H_0$ is the present Hubble parameter. We show that $\\rho(R)/\\rho_0=4(df/dR)^2/f^6$ and $p(R)/\\rho_0=-4/3(d^2f/dR^2)/f^5$ where $\\rho_0$ is the present density and $f(R)=1/\\sqrt{1+z(R)}$. This means that the pressure is in proportion to the curvature of the curve $f(R)$. For an accurate expression of $f(R)$ we propose a Pade-like form of $f(R)=1/\\sqrt{u}$ with $u\\equiv 1+\\sum\\limits_{n=1}^{N} u_nR^n$. For flat $\\Lambda$ model, the expansion up to $N=7$ has at most an error $< 0.2\\%$ for $z < 1.7$ and all value of $\\Lambda$. We also propose a general method to determine the equation of state of the universe which has $< N-1$ free parameters. ", "conclusions": "Now if x-matter consists of the scalar field $\\phi$ with the potential $V(\\phi)$, they are related to $\\rho_X$ and $p_X$ as \\ba \\left(\\frac{d\\phi}{dt}\\right)^2&=& \\rho_X+p_X,\\nonumber \\\\ V(\\phi)&=&\\frac{1}{2}(\\rho_X-p_X). \\nonumber \\ea Using $\\rho(R)$ and $p(R)$, we have \\ba &&\\phi-\\phi_0=\\frac{1}{\\sqrt{8\\pi G}}\\int_0^R\\sqrt{-3\\Omega_Mu+\\frac{2}{u}\\frac{d^2u}{dR^2}}dR\\equiv g(R),\\\\ &&V(\\phi)=\\frac{3H_0^2}{\\sqrt{16\\pi G}}\\left(2(\\frac{du}{dR})^2-\\Omega_Mu^3- \\frac{2u}{3}\\frac{d^2u}{dR^2}\\right) \\equiv h(R), \\ea where $\\phi_0$ is the present value of the scalar field . From Eq. (17) we have $R=g^{-1}(\\phi-\\phi_0).$ Then the potential is expressed as $V(\\phi)=h(g^{-1}(\\phi-\\phi_0))$. In section 3 we assumed that accurate $u_n$ for $n=1,\\ldots, N$ are obtained. We here show an example of the determination of $u_n$ for $n=1,\\ldots, 7$. We adopt the redshifts of 38 data with $z > 0.17$ from (\\cite{perl99}). We also adopt the relative error of $R$ for each data as $X_i$. Now let us assume that our universe obeys the $\\Lambda$ model with $\\Omega_M=0.3$. Then we know the theoretical value of $R_i^t$ for each $z_i$. To simulate the real observation, we set $R_i=R_i^t(1\\pm SX_i)$ where S is a scale factor. Let us assume that an accurate observation gives us 38 luminosity distances with $\\sim 0.1\\%$ accuracy so that the scale factor $S$ is chosen to make the relative error of $R$ for 38 data be $\\sim 0.1\\%$.\\footnote{This might be possible if the statistical error approaches the systematic error in, for example, the SNAP project. } We performed the likelihood analysis for this simulated data and obtained $u_2=0.2240, u_3=0.1505, u_4=0.0585$, $u_5=0.0187, u_6=0.0095$ and $ u_7=0.0053$ with $\\chi^2=36.16$ for 38-7 d.o.f. while theoretical values are $u_2=0.225, u_3=0.15, u_4=0.054375$, $u_5=0.016875, u_6=0.0068906$ and $ u_7=0.00225$. >From this $\\Omega_M=4/3u_2=0.29866$ or $\\Omega_M=2u_3=0.3010$ is obtained while $dp/d\\rho=0.0052$ at $u=1$. Assuming the more general equation of state with $w_2=w_3\\ldots=w_ 7=0$, we have $\\Omega_M= 0.31095, w_0= -1.01783$ and $w_1=-0.04768$. This suggests that we may confirm the $\\Lambda$ model if such accurate luminosity distances are available. We show in Fig. 2 the simulated data and the results of the likelihood analysis. Note that error bars are extended by a factor 100. The theoretical curve (the dashed line) and the observational curve (the solid line) are almost indistinguishable." }, "0011/astro-ph0011361_arXiv.txt": { "abstract": "Re-examination of extensive photometric data of TV~Col reveals evidence for a permanent positive superhump. Its period (6.4~h) is 16 percent longer than the orbital period and obeys the well known relation between superhump period excess and binary period. At 5.5-h, TV~Col has an orbital period longer than any known superhumping cataclysmic variable and, therefore, a mass ratio which might be outside the range at which superhumps can occur according to the current theory. We suggest several solutions for this problem. ", "introduction": "\\subsection{Permanent superhumps} Permanent superhump systems compose a new subclass of cataclysmic variables (CVs), whose existence was established only in the nineties. Systems that belong to this group have quasi-periodicities slightly shifted from their orbital periods, in addition to the binary periods themselves. Unlike SU~UMa systems (see Warner 1995 for a review on SU~UMa systems and CVs in general), which show this behaviour only during superoutbursts, in permanent superhump systems the phenomenon is observed during their normal brightness state. According to Osaki (1996), permanent superhumpers differ from other subclasses of non-magnetic CVs by their relatively short orbital periods and high mass transfer rates, resulting in accretion discs that are thermally stable but tidally unstable. Retter \\& Naylor (2000) provided observational support for this idea. The `$\\bf positive$ $\\bf superhump$', periodicity which is a few percent larger than the orbital period, is explained as the beat period between the binary motion and the precession of an eccentric disc in the apsidal plane. Periods slightly shorter than the orbital periods have also been seen in several systems. They are known as `$\\bf negative$ $\\bf superhumps$' (Patterson 1999). The observations show a roughly linear relation between the positive superhump period excess as a fraction of the binary period and the binary period (Stolz \\& Schoembs 1984). Negative superhumps seem to obey a similar rule (Patterson 1999). It has been suggested that negative superhumps are generated by the nodal precession of the accretion disc (Patterson et al. 1993; Patterson 1999); however, there are some theoretical difficulties with this idea (e.g. Murray \\& Armitage 1998). \\subsection{Periodicities in TV~Col} The periodicities detected so far in TV~Col and their common interpretations are (Motch 1981; Hutchings et al. 1981; Schrijver, Brinkman \\& van der Woerd 1987; Barrett, O'Donoghue \\& Warner 1988; Hellier, Mason \\& Mittaz 1991; Hellier 1993; Augusteijn et al. 1994): \\begin{itemize} \\item 4 day - the nodal precession of the accretion disc \\item 5.5 hr - the orbital period \\item 5.2 hr - the negative superhump (the beat between the orbital period and the nodal precession) \\item 32 min - the spin period \\end{itemize} There seems to be a strong connection between positive and negative superhumps. Light curves of many permanent superhumpers show both types of superhumps. In addition, period deficits in negative superhumps are about half period excesses in positive superhumps (Patterson 1999): $\\epsilon_{negative}$$\\approx$--0.5$\\epsilon_{positive}$, where $\\epsilon$=($P_{superhump}$--$P_{orbital})/P_{orbital}$. We, therefore, decided to look in available photometric data on TV~Col for positive superhumps, which would be predicted to have a period near 6.4 h. Here we report the finding of such a periodicity as initially announced by Retter \\& Hellier (2000). ", "conclusions": "The photometric data show evidence for a periodicity of 0.265 d in addition to the previously known periods. The repeatability of the peak in three independent datasets makes it 95\\% significant. In addition, the better the data are, the more the period stands out of the noise. Moreover, it has almost exactly the value predicted from the Stolz \\& Schoembs (1984) relation (updated by Patterson 1999) shown in Fig.~2. TV~Col has already been classified as a permanent superhump system because its 5.2-h period was interpreted as a negative superhump. In addition, the new period and the negative superhump obey the relation between the two types of superhumps (Section~1.2). Therefore, the new period is naturally interpreted as a positive superhump. \\begin{figure} \\centerline{\\epsfxsize=2.75in\\epsfbox{fig2.eps}} \\caption{The relation between superhump period excess (over the binary period) and binary period in superhump systems. Empty circles correspond to periods in the SU~UMa systems. Filled circles represent permanent superhumpers. The solid line represents the linear fit to the data. The two tilted dashed lines show the 1-$\\sigma$ error. TV~Col obeys the relation. The upper edge of the period gap (as defined by Diaz \\& Bruch 1997) is marked by the vertical long-dashed line.} \\end{figure} According to theory superhumps can appear only in CVs with small mass ratios -- q=$ M_{donor}/M_{compact}$$<$0.33. Hellier (1993) concluded, however, that q=0.62-0.93 from a spectroscopic analysis of the system, but this depended on an interpretation of the emission lines that may not be correct. Using the superhump excess, we find: q=0.95$\\bf \\pm$0.10 -- well above the 0.33 limit suggested by the hydrodynamic simulations, and consistent with the values estimated by Hellier. The mass ratio in TV~Col may thus be above the theoretical limit, perhaps due to its strong magnetic field. Alternatively, TV~Col may be an extreme system, with a very massive white dwarf near the Chandrasekhar mass (1.44$\\bf M_{\\odot}$), and / or an undermassive secondary star." }, "0011/astro-ph0011407_arXiv.txt": { "abstract": "We present photometry and high SNR spectroscopy in the classification region of V635~Cas, the optical counterpart to the transient X-ray pulsator 4U\\,0115+63, taken at a time when the circumstellar envelope had disappeared. V635~Cas is classified as a B0.2Ve star at a distance of $7-8$ kpc. We use the physical parameters derived from these observations and the orbit derived from X-ray observations to elaborate a model of the system based on the theory of decretion discs around Be stars. We show that the disc surrounding the Be star must be truncated by the tidal/resonant interaction with the neutron star and cannot be in a steady state. This explains many of the observed properties of 4U\\,0115+63. In particular, because of this effect, under normal circumstances, the neutron star cannot accrete from the disc, which explains the lack of regular Type~I outbursts from the source. ", "introduction": "The hard X-ray transient 4U\\,0115+63 (X\\,0115+634) is one of the best studied Be/X-ray binary systems (see Campana 1996; Negueruela et al. 1997, henceforth N97). More than 50 of these systems, in which a neutron star orbits a Be star in a moderately eccentric orbit, are known (see Negueruela 1998; Bildsten et al. 1997). The Be star is surrounded by a disc of relatively cool material, presumably ejected from the star due to causes unknown, but generally believed to be associated with fast rotation, magnetic fields and/or non-radial pulsations (Slettebak 1988). The presence of the disc gives rise to emission lines in the optical and infrared spectral regions and an excess in the infrared continuum radiation. The hard X-ray emission is due to the accretion of circumstellar material on to the neutron star companion. Due to their different geometries and the varying physical conditions in the circumstellar disc, Be/X-ray binaries can present very different states of X-ray activity (Stella et al. 1986). In quiescence, they display persistent low-luminosity ($L_{{\\rm x}} \\la 10^{36}$ erg s$^{-1}$) X-ray emission or no detectable emission at all. Occasionally, they show series of periodical (Type I) X-ray outbursts ($L_{{\\rm x}} \\approx 10^{36} - 10^{37}$ erg s$^{-1}$), separated by the orbital period of the neutron star. More rarely, they undergo giant (Type II) X-ray outbursts ($L_{{\\rm x}} \\ga 10^{37}$ erg s$^{-1}$), which do not clearly correlate with the orbital motion. Some systems only display persistent emission, but most of them show outbursts and are termed Be/X-ray transients. \\begin{figure*} \\begin{picture}(500,250) \\put(0,0){\\special{psfile=fig1.ps angle =0 hoffset=-100 voffset=-240 hscale=120 vscale=120}} \\end{picture} \\caption{The spectrum of V635~Cas in the classification region. Two exposures taken on November 14, 1997, with ISIS on the WHT equipped with the R1200B grating and the EEV10 camera have been combined for this figure. The comparison spectrum is that of the B0.2V standard $\\tau$ Sco. Both spectra have been divided by a spline fit to the continuum for normalisation and smoothed with a $\\sigma = 0.8$\\AA\\ Gaussian function for display.} \\label{fig:bluespec} \\end{figure*} The transient 4U\\,0115+63 was first reported in the {\\em Uhuru} satellite survey (Giacconi et al. 1972; Forman et al. 1978), though a search of the {\\em Vela 5B} data base revealed that the source had already been observed by this satellite since 1969 (Whitlock et al. 1989). Precise positional determinations by the {\\em SAS 3}, {\\em Ariel V} and {\\em HEAO-1} satellites (Cominsky et al. 1978; Johnston et al. 1978) were used to identify the system with a heavily reddened Be star with a visual magnitude $V \\approx 15.5$ (Johns et al. 1978; Hutchings \\& Crampton 1981), which was subsequently named V635~Cas (Khopolov et al. 1981). Rappaport et al. (1978) used {\\em SAS 3} timing observations to derive the orbital parameters of the binary system, which consists of a fast-rotating ($P_{{\\rm s}} = 3.6 \\: {\\rm s}$) neutron star in a relatively close ($P_{{\\rm orb}} = 24.3 \\: {\\rm d}$) and eccentric ($e = 0.34$) orbit around the Be star (see also Tamura et al. 1992). Due to the fast rotation of the neutron star, centrifugal inhibition of accretion prevents the onset of X-ray emission unless the ram pressure of accreted material reaches a relatively high value (Stella et al. 1986; N97). The system had only be known to display Type II activity until a short series of Type I outbursts was detected by BATSE and {\\em RXTE} in 1996 (Bildsten et al. 1997; Negueruela et al. 1998). The giant outbursts are associated with large amplitude brightenings of the optical and infrared magnitudes of the counterpart (N97). This is the first of two papers dedicated to providing a coherent picture of this system and understanding the implications of its unusual behaviour for the general class of Be/X-ray transients. Here we derive the astrophysical parameters of 4U\\,0115+63 and build a model for circumstellar disc around the Be star that allows us to understand the usual quiescent state of the system. In the second paper (Negueruela et al. 2000; henceforth Paper II), we will analyse the temporal evolution of this disc and investigate how its behaviour is connected with the X-ray activity of the source. ", "conclusions": "By observing the source during a disc-loss episode, we have been able to determine the stellar parameters of V635~Cas. The object is classified as a B0.2Ve star at a distance of $7-8$ kpc. Both the mass function and the estimated $v\\sin i$ indicate a moderate inclination for the orbital and equatorial planes. The derived distance implies that the source can radiate close to the Eddington luminosity for a neutron star during bright outbursts. With the newly determined parameters, we have constructed a model for 4U\\,0115+63. Based on the viscous decretion disc model for Be stars, we have numerically solved a criterion for tidal truncation and found that the disc surrounding V635~Cas must be truncated at a resonance radius depending on the viscosity parameter and cannot be in a steady state. Although we have adopted a particular disc model, our conclusion is robust as long as the outflow in the disc is subsonic, a hypothesis supported by several observational facts. Under normal conditions, the neutron star cannot accrete enough material to overcome the centrifugal barrier and switch on the X-ray emission." }, "0011/astro-ph0011541_arXiv.txt": { "abstract": "Violation of the axial symmetry of a magnetic field essentially modifies the physics of the plasma outflow in the magnetosphere of rotating objects. In comparison to the axisymmetric outflow, two new effects appear: more efficient magnetocentrifugal acceleration of the plasma along particular field lines and generation of MHD waves. Here, we use an ideal MHD approximation to study the dynamics of a cold wind in the nonaxisymmetric magnetosphere. We obtain a self-consistent analytical solution of the problem of cold plasma outflow from a slowly rotating star with a slightly nonaxisymmetric magnetic field using perturbation theory. In the axisymmetric (monopole-like) magnetic field, the first term in the expansion of the terminating energy of the plasma in powers of $\\Omega$ is proportional to $\\Omega^4$, where $\\Omega$ is the angular velocity of the central source. Violation of the axial symmetry of the magnetic field crucially changes this dependence. The first correction to the energy of the plasma becomes proportional to $\\Omega$. Efficient magnetocentrifugal acceleration occurs along the field lines curved initially in the direction of the rotation. I argue that all the necessary conditions for the efficient magnetocentrifugal acceleration of the plasma exist in the radio pulsar magnetosphere. We calculated the first correction of the rotational losses due to the generation of the MHD waves and analysed the plasma acceleration by these waves. ", "introduction": "The outflow of plasma from rotating magnetised objects is a widespread phenomenon in the Universe. It is observed in a wide range of astrophysical objects of different natures and scales. Some of these objects eject relativistic plasma. The Lorentz factor of plasma ejected by microquasars and AGN is of the order 10 (Mirabel \\& Rodrigues \\cite{mirabel}, Begelman et al. \\cite{bbr}, Pelletier et al. \\cite{pelletier}), although there is evidence that in Blazars the plasma is accelerated to the Lorentz factor $\\sim 10^6$ (Aharonian et al. \\cite{ah}). Radio pulsars also accelerate the plasma to the Lorentz factor $\\sim 10^6$ (Kennel \\& Coroniti \\cite{kennel}, Coroniti \\cite{coroniti}, Arons \\cite{arons96}). The problem of the acceleration of the plasma to such high energies remains one of the most important unsolved problems in modern relativistic astrophysics. The acceleration of plasma can occur in principle due to different dissipative as well as nondissipative processes. Hypothesis have been proposed about the singular behaviour of the relativistic plasma outflow at the ``light surface'' (\\cite{beskin83}) or at the light cylinder (Mestel \\& Shibata \\cite{ms}) of radio pulsars. These dissipative structures could provide the plasma acceleration. However, recent direct numerical calculations of the relativistic outflows in MHD approximation (Bogovalov \\cite{bog97}) and force-free approximation (Contpoulos et al. \\cite{janis}) have demonstrated that the relativistic plasma flow is actually regular everywhere. Another mechanism of acceleration due to dissipative processes was proposed by Michel (1982) and Coroniti (1990) in application to radio pulsars. Michel was the first who noticed that the number of charge carriers in the wind from radio pulsars is not enough to sustain the necessary return current in the wind with a striped magnetic field formed at the outflow from the oblique rotator. Coroniti (1990) considered the reconnection process in this magnetic field and came to the conclusion that it can provide the necessary acceleration of the winds from radio pulsars. Up to now this process was considered as the only possible mechanism explaining pulsar wind acceleration. However, recently Luybarskii \\& Kirk (2000) have revisited the process of wind acceleration due to magnetic field reconnection. They concluded that well before the winds from pulsars such as the Crab pulsar reach high Lorentz factors they will be terminated by the interstellar medium. A lot of efforts has been spent to obtain the plasma acceleration in frameworks of ideal MHD. There is a general opinion that in all astrophysical objects ejecting plasma, the magnetic field plays an essential role in the dynamics of the plasma. In particular, the magnetic field of a rotating object itself can provide the acceleration of the plasma due to a magnetocentrifugal (below MC) mechanism (Blandford \\& Payne \\cite{bp}). The ideal MHD outflows of plasma in different models and different approximations were investigated in a series of works (Michel \\cite{michel69}, Sulkanen \\& Lovelace \\cite{lovlace}, Ferreira \\cite{ferreira}, Vlahakis \\& Tsinganos \\cite{tsin}, Bogovalov \\& Tsinganos \\cite{bogtsin}, Krasnopolsky et al. \\cite{kb}, Ustyugova et al. \\cite{ustyug}). It follows from these studies that nonrelativistic plasma can be accelerated rather efficiently due to this mechanism. As far as relativistic plasma is concerned, the efficiency of the acceleration of this plasma appears insufficient for astrophysical applications. There were hopes that divergence of the magnetic flux tubes somewhere beyond the light cylinder will result in more efficient acceleration of the plasma due to the pressure gradient of the toroidal magnetic field (Begelman \\& Li \\cite{bl}, Takahashi \\& Shibata \\cite{takshib}). The only known mechanism for this magnetic flux tube divergence is magnetic self-collimation of the magnetised winds (Heyvaerts \\& Norman \\cite{heyverts}, Chiueh et al. \\cite{chuk}). This mechanism can, in principle, provide divergence of the field lines at the equatorial plane such that the poloidal field decreases with distance faster than $r^{-2}$. This divergence can take place in the limiting range of distances from the source, since finally the wind expands radially near the equator at $r\\rightarrow \\infty$ (Heyvaerts \\& Norman \\cite{heyverts}, Chiueh et al. \\cite{chuk}). Numerical self-consistent solution of the problem of steady-state plasma outflow in a model with the an initially monopole-like magnetic field shows that collimation really increases the efficiency of nonrelativistic plasma acceleration at the equator (Bogovalov \\& Tsinganos \\cite{bogtsin}). However, collimation of the relativistic plasma outflow appears negligibly small at large Lorentz factors (Bogovalov \\cite{bog97}, Bogovalov \\cite{bog00}). Therefore, the relativistic plasma is practically not accelerated by the pressure gradient of the toroidal magnetic field. This conclusion is valid not only for the monopole-like model. The acceleration of the wind due to the pressure gradient of the toroidal magnetic field occurs at the large distances compared to the dimension of the central source (Begelman \\& Li \\cite{bl}, Takahashi \\& Shibata \\cite{takshib}). The stationary wind from any source is monopole-like at these distances. Therefore, the poloidal magnetic field also becomes monopole-like, since the magnetic field is frozen into the plasma. Thus, the solutions obtained in the monopole-like model actually describe general properties of any axisymmetric wind at large distances from the source. The application of the monopole-like model is not limited by large distances. This model also gives upper limits on the efficiency of the MC acceleration of the plasma in any other axisymmetric magnetic field (dipolar for example). It is qualitatively clear that the slower the magnetic field decreases with distance, the more efficiently it accelerates the plasma due to a MC mechanism. The monopole-like magnetic field falls down as $r^{-2}$, while the dipolar magnetic field falls down faster than $r^{-2}$. Correspondingly, the plasma is expected to be more efficiently accelerated in the monopole-like magnetic field in compare with the acceleration in the dipolar one. This is why the results obtained in the monopole-like model rule out efficient MC acceleration of the plasma in any other more realistic axisymmetric model. Calculations performed by Contopoulos et al. (\\cite{janis}) confirm this conclusion. Thus, up to now all the attempts to propose an effective mechanism of relativistic plasma acceleration have not been successful. That is why the search for new possible mechanisms to explain relativistic plasma acceleration remains one of the most important problems of relativistic astrophysics. The disappointing conclusion regarding the low efficiency of the MC acceleration is valid only for axisymmetric magnetic fields, as previous studies considered this process only within the frameworks of axisymmetric models. However, the magnetic field of real astrophysical objects is not axisymmetric. Radio pulsars give us the brightest example of this relation. Therefore, it is important to know how the non axisymmetry of the magnetic field affects the process of MC acceleration. In this paper we try to answer this question. To clarify the role of the nonaxisymmetry of the magnetic field for the MC acceleration of plasma, a model with the initially monopole-like magnetic field is used in this paper. This model is often used to investigate the processes of plasma collimation and acceleration in the rotating magnetosphere (Michel \\cite{michel69}, Mestel \\& Selley \\cite{mestel}, Sakurai \\cite{sakurai}, Bogovalov \\cite{bog92}, Beskin et al. \\cite{vasia98}, Bogovalov \\& Tsinganos \\cite{bogtsin}, Tsinganos \\& Bogovalov \\cite{tsbog}). The word ``initially\" implies that the magnetic field of the non rotating star looks like the magnetic field of the ``magnetic monopole\". This model is very convenient since there are no closed field lines in the flow. It remarkably simplifies the analysis of the plasma acceleration. The main goal of this paper is to answer the question: how does the nonaxisymmetry of the magnetic field of a star affect the process of plasma acceleration? An attempt to solve the problem of the structure of the magnetosphere of an oblique rotator with a dipole magnetic field in mass-less approximation has been done by Beskin et al. (\\cite{beskin83}). The first step in the solution of this problem in the MHD approximation has been done in the work of Bogovalov (\\cite{bog99}). In this work, the problem of plasma flow in the magnetosphere of the oblique rotator with an initially split-monopole magnetic field was solved. The modulus of the magnetic field was axisymmetric and only the direction of the magnetic field varied with time and the azimuthal angle. It was found that the acceleration remains non efficient in this flow as well. In the present paper, we are interested in the affect of the nonaxisymmetry of the modulus of the magnetic field on the process of plasma acceleration. Plasma flow is described by a system of non linear equations in partial derivatives. Usually these equations can be integrated only numerically. There is one exception from this rule. Sometimes the problem can be solved self-consistently and analytically if we are interested in small corrections to the known solution. These corrections can arise if the known flow is perturbed slightly by additional small forces. In particular, slow rotation can be considered as a small perturbation of some known flow from the non rotating star. This approach was firstly successfully applied for the numerical self-consistent solution of the flow of the solar wind by Nurney and Suess (\\cite{ns}). Later Bogovalov (\\cite{bog92}) has used this approach for the fully analytical, self-consistent solution of the problem of the cold plasma outflow, relativistic as well as nonrelativistic, from a slowly rotating star. In recent years, this method has also been used to solve a range of problems by Beskin (see Beskin \\& Okamoto \\cite{bsknokmto}) and references therein). The idea of the present work is the following: if the nonaxisymmmetry of the magnetic field modifies the process of MC acceleration of the plasma, this modification can be seen at the level of the slow rotation of the object. Therefore, to answer the question: does the nonaxisymmetry of the magnetic field modify the process of acceleration of the plasma, we investigate the acceleration of plasma in the nonaxisymmetric magnetic field at slow rotation. This paper is organised as follows. In Sect. 2 we present ideal MHD equations defining the dynamics of cold relativistic plasma. A self-consistent analytical solution of the problem is given in Sec. 3. The solution in the wave zone is given in Sect. 4. The possible astrophysical implications of the results are discussed in Sec. 5. ", "conclusions": "" }, "0011/astro-ph0011294_arXiv.txt": { "abstract": " ", "introduction": "In the 19th century the American mathematician G.W. Hill devised a simple and useful approximation for the motion of the moon around the earth with perturbations by the sun. To most dynamical astronomers ``Hill's Problem''\\index{Hill's problem} still means a model for motions in the solar system in which two nearby bodies move in nearly circular orbits about another much larger body at a great distance. These lectures have, however, been motivated by a problem in stellar dynamics\\index{stellar dynamics}. Consider a star in a star cluster\\index{star cluster} which is itself in orbit about a galaxy (Figure \\ref{derivation}). The star, cluster and galaxy take the place of the moon, earth and sun, respectively. The potentials of the cluster and galaxy are not those of a point mass, and the galactic orbits of the star and cluster may be far from circular. Nevertheless Hill's problem is a good starting point, and it can be modified easily to accommodate the differences. In section 2 we outline a derivation of Hill's equations, and in section 3 we summarise the appropriate extensions. Stars gradually escape\\index{escape} from star clusters. This has been expected on theoretical grounds for many years, ever since a paper by Ambartsumian (1938). Recently, deep observations have confirmed this (e.g. Leon et al 2000), by revealing faint streams of stars around a number of the globular clusters of our Galaxy. Loosely speaking we can say that a star can only escape if its energy exceeds some critical energy. The energies of stars change slightly as a result of two-body gravitational encounters within clusters, though the time scale on which this happens (the {\\sl relaxation time\\index{relaxation time} scale}) is very long, of order $10^9$yr. But the orbital motions of stars within clusters have much smaller time scales of order $10^6$yr, and until recently it was thought that escaping stars would leave on a similar time scale. With this assumption, relaxation is the bottleneck, and so the escape time scale (e.g. the time taken for half the stars to escape) should vary with the relaxation time. Nowadays it is possible to simulate the evolution of modest-sized star clusters with $3\\times10^4$ or more members, and the predicted escape time scale can be checked empirically. Unfortunately the results contradict the theory (Figure {\\ref{collaborative}}). As these simulations\\index{$N$-body simulations} require considerable extrapolation in particle number $N$ to be applicable to real clusters (for which $N\\sim 10^6$) the error of the theory is serious. It turns out that the assumption of rapid escape is the main source of error (Fukushige \\& Heggie 2000, Baumgardt 2000a,b). In fact some stars above the escape energy never escape\\index{escape} (unless some other dynamical process comes into play), and others take much longer to escape than had been generally thought. \\begin{figure} [htbp] \\centering \\includegraphics[width=10cm,clip,trim=0 0 0 0,angle=-90]{th.eps} \\caption{Results of numerical experiments (Aarseth \\& Heggie, unpublished) on the escape of stars from star clusters\\index{star cluster}. The time for half the stars to escape is plotted against the original number $N$ of stars in the simulation. Points are averages over several simulations\\index{$N$-body simulations} at each $N$, except the largest value. The continuous line shows the prediction of theory, i.e. proportional to the relaxation time\\index{relaxation time} (see text), and the dashed line is an empirical fit.} \\label{collaborative} \\end{figure} With this motivation, the remaining sections of these lectures are devoted to the dynamics of escape. Section 4 analyses the very definition of escape, which is not as straightforward as in more familiar situations. The last two sections show some ways in which the computation of the escape rate can be approached. The main result of section 5 concerns the way in which the time scale of escape depends on the energy, and outlines how this resolves the problem of Figure {\\ref{collaborative}}. Much more difficult, from a theoretical point of view, is determining the {\\sl distribution} of escape times, and some relevant ideas are introduced in Section 6. ", "conclusions": "" }, "0011/astro-ph0011206_arXiv.txt": { "abstract": "After decades of one-dimensional nucleosynthesis calculations, the growth of computational resources has meanwhile reached a level, which for the first time allows astrophysicists to consider performing routinely realistic multidimensional nucleosynthesis calculations in explosive and, to some extent, also in non-explosive environments. In the present contribution we attempt to give a short overview of the physical and numerical problems which are encountered in these simulations. In addition, we assess the accuracy that can be currently achieved in the computation of nucleosynthetic yields, using multidimensional simulations of core collapse supernovae as an example. ", "introduction": " ", "conclusions": "" }, "0011/astro-ph0011030_arXiv.txt": { "abstract": "Associated absorption lines (AALs) are valuable probes of the gaseous environments near quasars. Here we discuss high-resolution (6.7 \\kms ) spectra of the AALs in the radio-loud quasar 3C 191 (redshift $z=1.956$). The measured AALs have ionizations ranging from \\ion{Mg}{1} to \\ion{N}{5}, and multi-component profiles that are blueshifted by $\\sim$400 to $\\sim$1400~\\kms\\ relative to the quasar's broad emission lines. These data yield the following new results. 1) The strengths of excited-state \\ion{Si}{2}$^*$ AALs indicate a density of $\\sim$300 \\pcc\\ in the Si$^+$ gas. 2) If the gas is photoionized, this density implies a distance of $\\sim$28 kpc from the quasar. Several arguments suggest that all of the lines form at approximately this distance. 3) The characteristic flow time from the quasar is thus $\\sim$$3\\times 10^7$ yr. 4) Strong \\ion{Mg}{1} AALs identify neutral gas with very low ionization parameter and high density. We estimate $n_H\\ga 5\\times 10^4$ \\pcc\\ in this region, compared to $\\sim$15 \\pcc\\ where the \\ion{N}{5} lines form. 5) The total column density is $N_{\\rm H}\\la 4\\times 10^{18}$ \\cmsq\\ in the neutral gas and $N_{\\rm H}\\sim 2\\times 10^{20}$ \\cmsq\\ in the moderately ionized regions. These column densities are consistent with 3C 191's strong soft X-ray flux and the implied absence of soft X-ray absorption. 6) The total mass in the AAL outflow is $M\\sim 2\\times 10^9$ \\Msun , assuming a global covering factor (as viewed from the quasar) of $\\sim$10\\% . 7) The absorbing gas only partially covers the background light source(s) along our line(s) of sight, requiring absorption in small clouds or filaments $<$0.01 pc across. The ratio $N_H/n_H$ implies that the clouds have radial (line-of-sight) thicknesses $\\la$0.2 pc. These properties might characterize a sub-class of AALs that are physically related to quasars but form at large distances. We propose a model for the absorber in which pockets of dense neutral gas are surrounded by larger clouds of generally lower density and higher ionization. This outflowing material might be leftover from a blowout associated with a nuclear starburst, the onset of quasar activity or a past broad absorption line (BAL) wind phase. ", "introduction": "Associated absorption lines (AALs) are important diagnostics of the gaseous environments of quasars and active galactic nuclei (AGNs). The lines are defined empirically as having velocity widths less than a few hundred \\kms\\ and absorption redshifts, $z_a$, within a few thousand \\kms\\ of the quasar's emission-line redshift, $z_e$ (Weymann \\etal 1979, Foltz \\etal 1986). The requirement for \\zaz\\ makes AALs more likely to be physically related to the quasars than are the other narrow absorption lines (at \\zllz ) in quasar spectra (see Rauch 1998 for a review of the unrelated \\zllz\\ systems). The narrow velocity widths further distinguish AALs from the class of broad absorption lines (BALs), whose widths and maximum blueshifted velocities both typically exceed 10,000 \\kms\\ (Weymann \\etal 1991). BALs clearly form in high-velocity winds from quasar engines (e.g. Turnshek 1988), but AALs can form potentially in a variety of environments --- ranging from energetic outflows like the BALs to relatively quiescent gas at large galactic or inter-galactic distances (see Hamann \\& Brandt 2000 for a general review, also Tripp \\etal 1996, Hamann \\etal 1997a, Barlow \\& Sargent 1997, Barlow, Hamann \\& Sargent 1997). More work is needed to locate individual AAL absorbers, quantify their kinematic and physical properties, and understand the role of the AGNs and/or host galaxies in providing the source of their material and kinetic energy. One interesting property is that, among radio-loud quasars, AALs appear more frequently and with greater strength in sources with ``steep'' radio spectra and/or lobe-dominated radio morphologies (Wills \\etal 1995, Barthel \\etal 1997, Richards \\etal 2000, also Brotherton \\etal 1998 and references therein). This weak correlation is usually attributed to an orientation effect, whereby AAL regions reside preferentially near a disk or torus that is aligned perpendicular to the radio jet axis (but see Richards \\etal 2000 for an alternative interpretation). However, it is not clear if this (possible) AAL geometry has its origins on small scales related to the inner black hole/accretion disk, or on much larger scales related to the host galaxy. It is also not clear if a disk-like geometry applies as well to the AALs in radio-quiet sources. 3C 191 (Q0802+103, $z_e = 1.956$) is a radio-loud quasar having both strong AALs (Burbidge, Lynds \\& Burbidge 1966, Stockton \\& Lynds 1966, Williams \\etal 1975, Anderson \\etal 1987) and a bipolar, lobe-dominated radio structure (Akujor et al. 1994). It therefore follows the AAL--radio morphology correlation noted above. 3C 191 also provides a rare opportunity to define the distance between the quasar and the absorbing gas because several of its AALs arise from excited energy states, e.g. \\ion{C}{2}$^*$ \\lam 1336 and \\ion{Si}{2}$^*$ \\lam 1265,1533 (Bahcall, Sargent \\& Schmidt 1967, Williams \\etal 1975). The strengths of the excited-state lines, compared to their resonant counterparts, \\ion{C}{2} \\lam 1335, \\ion{Si}{2} \\lam 1260,1527, provide measures of the gas density needed to populate the upper levels. The density in turn constrains the absorber's distance from the quasar, with the reasonable assumption that the gas is in photoioization equilibrium with the quasar radiation field. Williams et al. (1975) already estimated a density of $n_e\\sim 1000$ \\pcc\\ and a radial distance of $R\\sim 10$ kpc for the AAL region in 3C 191. We reobserved 3C 191 with higher spectral resolution and wider wavelength coverage to 1) obtain more reliable densities as a function of velocity, 2) search for \\ion{Fe}{2}$^*$ AALs, which might be revealing of a much higher density environment (Wampler \\etal 1995, Halpern \\etal 1996), 3) revisit the question of this absorber's origin and location, and 4) obtain better constraints on the AAL region dynamics, abundances and overall physical structure. Sections 2 and 3 below describe the observations and results. Section 4 provides measurements and analyses of the AALs. Section 5 draws further inferences and discusses physical models. Throughout this paper we define solar abundances by the meteoritic results in Grevesse \\& Anders (1989), and we use atomic transition data from the compilation by Verner, Verner \\& Ferland (1996). ", "conclusions": "\\subsection{Implications for the Structure and Dynamics} The results in \\S3 and \\S4 lead to following important conclusions regarding the AAL environment. 1) We have already noted that the absorber has a complex velocity structure that appears qualitatively similar in all lines and ions (\\S4.1). Stronger AALs, e.g. of higher ionization, are more smoothly distributed in velocity, but it seems clear that all of the lines trace the same overall physical structure. We conclude that the radial distance $R\\approx 28$ kpc derived for the Si$^+$ zone (\\S4.6) should be roughly characteristic of the entire AAL region (see also item 4 below). 2) The AAL gas appears to be outflowing from the quasar at velocities from $\\sim$400 to $\\sim$1400 \\kms\\ (\\S4.1). If the gas is not accelerating (or decelerating), then the time scale for this outflow reaching radius $R$ is \\begin{equation} t\\ \\approx\\ 3\\times 10^7\\ \\left({{R}\\over{28\\, {\\rm kpc}}}\\right) \\left({{1000\\, {\\rm km\\, s}^{-1}}\\over{v}}\\right)~~{\\rm yr} \\end{equation} where $v$ is a characteristic velocity. The overall AAL region will acquire a radial thickness during this expansion because different gas components move at different speeds. In 3C 191, the radial thickness, $\\Delta R$, could be comparable to the radius, $R\\approx 28$ kpc, because the velocity dispersion implied by the line widths, $\\Delta v \\approx 1000$ \\kms , is similar to the average flow speed. 3) The significant presence of \\ion{Mg}{1}, compared to \\ion{Mg}{2}, identifies a neutral gas component whose survival cannot be attributed to radiative shielding downstream from an \\ion{H}{2}--\\ion{H}{1} recombination front (\\S4.4). The \\ion{Mg}{1} region must have a low ionization parameter and therefore a high density. If we adopt $n_H\\approx 300$ \\pcc\\ and $\\log U\\approx -2.8$ for the \\ion{Si}{2} region (\\S4.6), then simple scaling based on $\\log U\\la -5$ in the neutral gas (\\S4.4) implies $n_H\\ga 5\\times 10^4$ \\pcc\\ in that region. In contrast, $\\log U\\sim -1.5$ in the \\ion{N}{5} zone indicates $n_H\\sim 15$ \\pcc\\ there. 4) The \\ion{Si}{2} \\lam 1527 and \\ion{Si}{2}$^*$ \\lam 1533 line profiles are surprisingly similar (Fig. 2) given the wide range of densities present. The nearly constant line ratio indicates a nearly constant density (within a factor of $\\sim$2) at all velocities. This result could be caused by a selection bias: the absorber may have a wide range of densities at each velocity, but the specific ionization requirements of Si$^+$ might persistently lead to just a narrow range in $U$ and therefore $n_{\\rm H}$ controlling the \\ion{Si}{2} and \\ion{Si}{2}$^*$ line strengths. In any case, the similar $n_{\\rm H}$ values inferred for different \\ion{Si}{2} velocity components support the argument (item 1 above) that the AAL region does not span a wide range in radial distance from the quasar. 5) Partial line-of-sight coverage of the quasar emission sources (\\S4.2) implies that the AAL clouds have characteristic sizes similar to or smaller than the projected area of the emitters. The \\Lya\\ and \\ion{C}{4} absorption lines sit atop strong BELs (Fig. 2); they might fully cover the continuum source while partly covering the larger BEL region. The characteristic size of these absorbers might therefore be as large as 0.1 -- 1 pc (i.e. the size of the BEL region, e.g. Peterson 1993, Kaspi \\etal 2000). However, the partial coverage inferred from the \\ion{Mg}{2}, \\ion{Al}{3} and \\ion{Fe}{2} lines must involve the continuum source, which requires absorber size scales $<$0.01 pc (for standard accretion disk models of the continuum emission, Netzer 1992). 6) Another constraint on the size scales comes from the ratio $N_H/n_H$. If the absorbing gas completely fills the volume it encompasses, then the radial thickness of the entire absorber would be of order, \\begin{equation} \\Delta R\\ \\approx\\ 0.2\\, \\left({{N_H}\\over{2\\times 10^{20}\\, {\\rm cm}^{-2}}}\\right) \\left({{300\\, {\\rm cm}^{-3}}\\over{n_H}}\\right) ~~{\\rm pc} \\end{equation} using parameters from \\S4.5 and \\S4.6. However, if the absorber is composed of discrete clouds that fill only part of the encompassed volume, then the overall AAL region could have a much greater radial thickness (item 2 above) while the individuals clouds are small compared to $\\Delta R$ in Equation 11. 7) The mass of the AAL region, $M$, depends on its radial distance, total column density and {\\it global} covering factor, $Q\\equiv \\Omega /4\\pi$ (where $\\Omega$ is the solid angle subtended by the absorber as ``seen'' from the central quasar). The value of $Q$ is not known, but it is not likely to exceed the detection frequency of AALs among radio-loud quasars\\footnote{The detection frequency sets an approximate upper limit on $Q$ because some fraction of the systems counted as AALs will have a different physical origin than the absorber in 3C 191 (\\S1).}, e.g. $\\sim$30\\% (G. Richards, private communication). If we let $Q_{0.1}$ represent the covering factor relative to $Q=0.1$, the total mass is given by $$M\\ \\approx\\ 2\\times 10^{9} ~ Q_{0.1} \\left({{\\mu_H}\\over{1.4}}\\right) \\left({{N_H}\\over{2\\times 10^{20}\\, {\\rm cm}^{-2}}}\\right) $$ \\begin{equation} \\times\\ \\left({{R}\\over{28\\, {\\rm kpc}}}\\right)^2 ~{\\rm M}_{\\odot} \\end{equation} where $\\mu_H$ is the mean molecular weight per H particle. The total kinetic energy in the AAL outflow is therefore, \\begin{equation} K\\ \\approx\\ 2\\times 10^{58}\\,\\left({{M}\\over{3\\times 10^9\\, \\Msun}}\\right) \\left({{v}\\over{1000\\, {\\rm km\\, s}^{-1}}}\\right)^2 ~ {\\rm ergs} \\end{equation} For comparison, the much faster ($v\\sim 10^4$ \\kms ) but smaller-scale ($R\\sim 0.1$ pc) BAL outflows observed in other quasars are believed to contain total masses (at any instant) of $\\sim$1--10 \\Msun\\ for $Q_{0.1}\\sim 1$ to 3. Over a quasar's lifetime, say 10$^8$ yr (Haehnelt, Natarajan \\& Rees 1998), a BAL wind could eject a total of $\\sim$10$^7$--10$^8$ \\Msun\\ with $K \\approx\\ 10^{58}$--10$^{59}$ ergs (Hamann \\& Brandt 2000). 8) Emission lines from the AAL gas could be quite strong, depending on the actual values of $U$ or $N_{\\rm H}$. For example, we calculated emission line strengths for a photoionized plasma consistent with the \\ion{Si}{2} AAL region described above, namely, having solar abundances, an incident spectrum as defined in \\S4.4, and the following physical paramters: $R\\approx 28$ kpc, $\\log U\\approx -2.8$, $N_{\\rm H} = 2\\times 10^{20}$ \\cmsq , and $n_H\\approx 300$ \\pcc . The strongest predicted emission lines within our wavelength coverage for 3C 191 are \\Lya\\ and \\ion{Mg}{2} \\lam 2799, with rest-frame equivalent widths of $\\sim$44$Q_{0.1}$ \\AA\\ and $\\sim$4$Q_{0.1}$ \\AA . These results should be compared to the rest-frame equivalent widths of the measured BELs, e.g. $\\sim$120 \\AA\\ in \\Lya\\ and $\\sim$17 \\AA\\ in \\ion{Mg}{2} (as estimated without the superposed AALs). The predicted emission lines could therefore be present but ``hidden'' in the measured BELs. Clearly, however, $Q_{0.1}\\sim 1$ is close to an upper limit by this test. Another constraint is that $Q_{0.1}>1$ could lead to too much line emission ``filling in'' the bottoms of the AAL troughs (depending on the global line-of-sight velocity distribution of the emitting gas; see Hamann, Korista \\& Morris 1993 and Hamann \\& Korista 1996 for similar arguments related to BALs). Future observations at longer wavelengths might provide more stringent upper limits on $Q_{0.1}$. In particular, our calculations predict that the strongest lines in the rest-frame near-UV/visible should be H$\\alpha$, H$\\beta$, the \\ion{O}{2} \\lam 3727 doublet, and \\ion{O}{3} \\lam 5007, with rest equivalent widths of $\\sim$$55Q_{0.1}$ \\AA , $\\sim$11$Q_{0.1}$ \\AA , $\\sim$22$Q_{0.1}$ \\AA , and $\\sim$88$Q_{0.1}$ \\AA , respectively. \\subsection{Toward a Physical Model} Figure 6 shows a highly schematic model of the AAL region, wherein pockets of dense neutral gas are surrounded by a diffuse, spatially distributed medium of generally higher ionization. The diffuse clouds contain most of the total column density (\\S4.5). Their greater size and/or greater numbers lead to more complete coverage in both velocity and projected area (for example, in the \\ion{C}{4} doublet), compared to the lower-ionization gas (e.g. \\ion{Mg}{1} \\lam 2853). Note that the extended regions cannot have exclusively high ionization levels because the measured line-of-sight coverage fractions do not correlate simply with ionization. For example, low ionization lines can have either high (e.g. \\ion{C}{2} \\lam 1334, \\ion{Si}{2} \\lam 1260) or low (\\ion{Fe}{2}) coverage fractions (\\S4.2). The amount of coverage in both space and velocity must depend at least partly on the line's oscillator strength. In other words, the coverage fraction scales with the line optical depth. Stronger (more optically thick) lines have greater contributions from the diffuse extended gas, resulting in greater coverage, whereas weak lines sample mainly the higher column density material in more compact regions. A major concern with any cloud model is the cloud survival. The lifetime of a cloud without pressure confinement is of order the sound-crossing time. For a nominal temperature of $10^4$ K and a maximum cloud size of $\\Delta R \\la\\ 0.2$ pc (Eqn. 11), the cloud survival time, $\\la$$2\\times 10^4$ yr, is much less than the characteristic flow time, $t\\approx\\ 3\\times 10^7$ yr (Eqn 10). Therefore pressure confinement appears necessary. The problem of understanding this confinement has plagued cloud models of both the BAL and BEL regions of AGNs. Possible solutions include external pressure from a magnetic field or a surrounding hot, low-density (and transparent) plasma (Weymann, Turnshek \\& Christiansen 1985, Arav, Li \\& Begelman 1994, Emmering, Blandford \\& Shlosman 1992, DeKool 1997, Feldmeier \\etal 1997). The key remaining question is, what provides the source of material and kinetic energy for the AAL outflow? The flow time ($\\sim$$3\\times 10^7$ yr) is comparable to predicted quasar lifetimes (Terlevich \\& Boyle 1993, Haehnelt et al. 1998). It therefore seems likely that the AAL gas originated much nearer the quasar, perhaps coincident with the onset of quasar activity. We have already noted (\\S5.1, item 7) that the kinetic energy in this AAL region is comparable to the typical energy in BAL winds (integrated over a quasar's lifetime). Therefore, quasars are capable of driving winds with this total energy. In addition, most current models of BAL winds have them preferentially located near the plane of the accretion disk (Murray \\& Chiang 1995, Emmering \\etal 1992, Wills, Brandt \\& Laor 1999 and references therein). This geometry is reminiscent of the equatorial structure (tentatively) inferred for AAL regions in radio-loud quasars (\\S1). However, it is unlikely that the AAL outflow in 3C 191 is simply an extended remnant of a BAL wind because 1) the terminal velocity of a BAL wind should be of order 10$^4$ \\kms\\ instead of 1000 \\kms , and 2) the total mass in AAL gas is at least an order of magnitude larger than expected for BAL winds (item 7 in \\S5.1). To produce the observed AALs, a high-velocity BAL-like wind would have to be decelerated by interaction with ambient galactic material and then, probably, entrain some of that material (to add mass) along the way. An alternative possibility is that the AALs form in gas that was expelled by stellar processes, e.g. in a galactic ``superwind'' as observed in low-redshift starburst galaxies (Heckman, Armus \\& Miley 1990, Heckman \\etal 2000). The sizes, masses, velocities, etc. inferred for superwinds in luminous starbursts are consistent with our estimates of these quantities in 3C 191. The superwinds also contain cool dense clouds (giving rise to \\ion{Na}{1} absorption lines) embedded in a hot ($\\sim$10$^7$ K) X-ray emitting plasma (see also Heckman \\etal 1996 and references therein). If a superwind model does apply to 3C 191, the characteristic flow time of the AAL gas ($\\sim$$3\\times 10^7$ yr) might represent the time elapsed since the starburst episode. No matter what scenario accounts for the AALs in 3C 191, it is important to keep in mind that AALs in different objects can probe very different physical phenomena. For example, AALs in other quasars often have higher blueshifted velocities than those in 3C 191. Some narrow absorption lines have blueshifts above the arbitrary 5000 \\kms\\ AAL threshold, even though there is strong evidence for their being intrinsic to the quasar environments (Hamann \\etal 1997a and 1997b, Barlow \\& Sargent 1997, Barlow, Hamann \\& Sargent 1997, Richards 2000). A galactic superwind certainly cannot explain these high-velocity absorbers. Most AAL systems also do not have low-ionization lines like 3C 191 (e.g. Junkkarinen, Hewitt \\& Burbidge 1991, Hamann 1997). It is possible that low ionization AALs, which allow us to locate the absorber via \\ion{Si}{2}$^*$, select in favor of large absorber--quasar distances. In particular, all of the known AAL systems with these excited-state lines have distances $\\ga$10 kpc (e.g. Barlow \\etal 1997, Tripp, Lu \\& Savage 1996, Morris \\etal 1986, Sargent, Boksenberg \\& Young 1982). Other AALs are known to form much closer to the quasars, possibly within a few pc in outflows similar to the BALs (Hamann \\etal 1997a and 1997b, Barlow \\& Sargent 1997, Barlow \\etal 1997). Given this diversity, it is interesting to note that 3C~191 does not follow the trend identified by Brandt \\etal (2000) for small X-ray to UV continuum flux ratios accompanying strong \\ion{C}{4} absorption equivalent widths (\\S4.5). That correlation nominally points to a relationship between the strength of the AALs and the strength of continuous (bound-free) absorption in X-rays. BAL quasars are at one extreme in this relationship --- having both strong UV lines and strong absorption in X-rays (see also Green \\& Mathur 1996, Gallagher \\etal 1999). 3C 191 might contain a different class of absorber (e.g. much farther from the active nucleus) than the majority of sources discussed by Brandt \\etal (2000)." }, "0011/astro-ph0011389_arXiv.txt": { "abstract": "We present a comprehensive study of the spectrum of the narrow-line Seyfert 1 galaxy \\rej, summarizing the information obtained from the optical to X-rays with observations from the William Herschel 4.2m Telescope (WHT), the Hubble Space Telescope (\\hst), the Extreme UltraViolet Explorer (\\euve), \\rosat, \\asca\\ and \\sax. The \\sax\\ spectra reveal a soft component which is well-represented by two blackbodies with $kT_{\\rm eff}$=60~\\eV\\ and 160~\\eV, mimicking that expected from a hot, optically-thick accretion disc around a low-mass black hole. This is borne out by our modeling of the optical to X-ray nuclear continuum, which constrains the physical parameters of a NLS1 for the first time. The models demonstrate that \\rej\\ is likely to be a system with a nearly edge-on accretion disk (60 to 75$^\\circ$ from the disk axis), accreting at nearly Eddington rates (0.3 to 0.7 $L_{\\rm Edd}$) onto a low mass black hole (M$_{bh}\\sim$2 to 10$\\times 10^6$ M$_{\\odot}$). This is consistent with the hypothesis that NLS1s are Seyfert-scale analogies of Galactic Black Hole Candidates. The unusually high temperature of the big blue bump reveals a flat power-law like continuum in the optical/UV which is consistent with an extrapolation to the hard X-ray power-law, and which we speculate may be similar to the continuum component observed in BL Lac objects in their quiescent periods. From the \\sax\\ and \\asca\\ data, we find that the slope of the hard X-ray power-law depends very much on the form of the soft component which is assumed. For our best-fitting models, it lies somewhere between $\\alpha$=0.7 and 1.3 and thus may not be significantly softer than AGN in general. ", "introduction": "The narrow-line Seyfert 1 (NLS1) type of AGN has proven to be a valuable and a fascinating resource for the study of the class as a whole. They have very strong soft X-ray excesses \\citep{puc92,bol96}, are often highly variable \\citep{bra96,bol97} and, in some cases, the optical/UV big blue bump (BBB) component is so hot that it is shifted into the UV/EUV regime, leaving a bare, power-law-like continuum component in the optical \\citep{puc95}. One of the earliest hypotheses put forward to explain the strong and variable ultra-soft X-ray excesses, was a high mass accretion rate onto a relatively low-mass black hole \\citep{pou95}. This was originally proposed for \\rej, by analogy with the properties of Galactic Black Hole Candidates (GBHCs). An alternative model was that such systems were geometrically-thick accretion disks (ADs) viewed face-on, lying co-planar with a flattened broad line region [BLR; \\citet{puc92}]. In our initial study of \\rej\\ \\citep{puc95}, we showed that the overall IR to X-ray spectrum compared well with the combination of a geometrically-thin, optically-thick AD and an underlying power-law with a slope, $\\alpha$=1.3 ($\\alpha$ is defined throughout such that $F_\\nu\\propto\\nu^{-\\alpha}$). The Kerr black hole mass was 7$\\times10^5$ M$_\\odot$, the accretion rate was 0.073 M$_\\odot$ yr$^{-1}$ and the disk was viewed at an angle of 60$^\\circ$ from its axis. However, this was for illustrative purposes only and no constraints were placed on these parameters. Thus, while this might have suggested the presence of a low-mass black hole for \\rej, consistent with the GBHC analogy, solutions at higher black hole mass could not be ruled out. With the launch of {\\sl Beppo-SAX} and the inclusion of \\rej\\ in the NLS1 core program, came the opportunity to place meaningful constraints on black hole mass (M$_{bh}$), accretion rate (\\.M) and inclination. \\rej\\ was also due to be observed with {\\sl HST} within a few months of {\\sl Beppo-SAX}, and an optical spectrum was also scheduled in {\\sl WHT} service time within a few weeks. Thus a quasi-simultaneous spectrum, providing the most complete coverage possible from the optical to mid X-rays was available. Simultaneity is important for a NLS1 due to the variable nature of the class, although \\rej, contrary to its counterparts, has shown remarkable stability. Thus these data have provided the best opportunity yet to fit AD models and provide constraints on the defining parameters of an NLS1. Using a combination of a power-law and an optically-thick, geometrically-thin Kerr AD model, we present fits to the observed spectrum of \\rej, measuring M$_{bh}$ and \\.M and constraining the inclination for the first time, and discuss the results in the context of the GBHC model. The improved UV and optical data presented here have also allowed us to re-visit the issue of what is producing the optical/UV continuum in this object. Our previous work had suggested a power-law-like continuum rising towards the red, which was not due to the host galaxy or the BBB. We have separated out the galaxy component from the spatially-resolved optical spectrum obtained and, when combined with the {\\sl HST} data, examined the pure nuclear component with a greater degree of clarity. The results and their implications for the production of the optical continuum in AGN are also discussed in this paper. ", "conclusions": "We have presented a comprehensive study of the narrow-line Seyfert 1 galaxy \\rej, spanning the optical to X-rays and six years of observations. Focusing on quasi-simultaneous \\sax, \\hst, \\euve\\ and optical data taken in early 1997, we have {\\sl (1)} measured the soft and hard X-ray components simultaneously for the first time using \\sax; {\\sl (2)} placed tight constraints on the form of the big blue bump using \\hst, \\euve\\ and \\sax; {\\sl (3)} separated out the nuclear component from the host galaxy in the optical; and {\\sl (4)} fitted the optical to X-ray nuclear continuum with the combination of an accretion disk and an underlying power-law, to constrain the black hole mass, accretion rate and inclination of the accretion disk in \\rej. Coverage of the widest possible range in X-rays, particularly at soft energies, is of prime importance for NLS1s, because of their strong ultra-soft components. These \\sax\\ data, which cover the 0.1 to 10~\\keV\\ range, provide significant constraints on the shape, and thus the physical nature, of the X-ray spectrum in \\rej. Our analysis of the \\sax\\ spectrum reveals a significant hardening of the X-ray spectrum above $\\sim$3~\\keV\\ and that two blackbodies, with $kT_{\\rm eff}$=60~\\eV\\ and 160~\\eV\\ best represent the ultra-soft X-ray component, consistent with multi-temperature emission from an optically-thick accretion disc. We do not find any compelling evidence to suggest that the hard X-ray power-law in \\rej\\ is significantly softer than that of non-NLS1 AGN. There is the suggestion of a low-energy flattening in the \\sax\\ data, which was also observed in the earlier \\rosat\\ PSPC spectrum. This in turn implies a high temperature for the inner edge of the accretion disk, and thus a low black hole mass (which is also borne out by the modeling). The results of fitting the optical to X-ray nuclear continuum have shown that the data prefer relatively high mass accretion rates ($\\sim$0.3 to 0.7 $\\dot M_{\\rm Edd}$) onto a low mass black hole (M$_{bh}\\sim$2-10$\\times 10^6$ M$_{\\odot}$) at high inclination angles (ie. preferably edge-on, $\\sim 60-70^\\circ$ away from the axis of the disc). The very low intrinsic columns implied by the X-ray fits suggest that any molecular torus must either be optically thin, geometrically thin, or lie out of the plane of the accretion disk so that it does not obscure our line of sight. The optical to X-ray nuclear continuum {\\sl requires} the presence of a flat, power-law-like component in the optical/UV. This is consistent with an extrapolation to the hard X-ray spectrum, although an independent origin for the hard X-ray power-law cannot be ruled out. The existence of such a component has been proposed before, but we believe that this is the first direct observation, by virtue of the unusually hot big blue bump in \\rej. It appears to be stable and unpolarized and we speculate that it may be a BL~Lac-type component which is currently in the quiescent stage of its duty-cycle." }, "0011/astro-ph0011340_arXiv.txt": { "abstract": "We present contemporary infrared and optical spectroscopic observations of the type~IIn SN~1998S covering the period between 3 and 127 days after discovery. During the first week the spectra are characterised by prominent broad H, He and C~III/N~III emission lines with narrow peaks superimposed on a very blue continuum (T$\\sim$24000~K). In the following two weeks the C~III/N~III emission vanished, together with the broad emission components of the H and He lines. Broad, blueshifted absorption components appeared in the spectra. The temperature of the continuum also dropped to $\\sim$14000~K. By the end of the first month the spectrum comprised broad blueshifted absorptions in H, He, Si~II, Fe~II and Sc~II. By day~44, broad emission components in H and He reappeared in the spectra. These persisted to as late as $\\sim$100-130 days, becoming increasingly asymmetric. We agree with Leonard {\\it et al.} (2000) that the broad emission lines indicate interaction between the ejecta and circumstellar material (CSM) emitted by the progenitor. We also agree that the progenitor of SN 1998S appears to have gone through at least two phases of mass loss, giving rise to two CSM zones. Examination of the spectra indicates that the inner zone extended to $\\leq$ 90~AU, while the outer CSM extended from 185~AU to over 1800~AU. We also present high resolution spectra obtained at 17 and 36 days. These spectra exhibit narrow P~Cygni H~I and He~I lines superimposed on shallower, broader absorption components. Narrow lines of [N~II], [O~III], [Ne~III] and [Fe~III] are also seen. We attribute the narrow lines to recombination and heating following ionisation of the outer CSM shell by the UV/X-ray flash at shock breakout. Using these lines we show that the outer CSM had a velocity of 40--50~km/s. Assuming a constant velocity, we can infer that the outer CSM wind commenced more than 170 years ago, and ceased about 20~years ago, while the inner CSM wind may have commenced less than 9 years ago. During the era of the outer CSM wind the outflow from the progenitor was high - at least $\\sim2\\times 10^{-5}$M$_{\\odot}$~yr$^{-1}$. This corresponds to a mass loss of at least $\\sim$0.003~$M_{\\odot}$, suggesting a massive progenitor. The shallower, broader absorption is of width $\\sim$350~km/s and may have arisen from a component of the outer CSM shell produced when the progenitor was going through a later blue supergiant phase. Alternatively, it may have been produced by the acceleration of the outer CSM by the radiation pressure of the UV precursor. We also describe and model first overtone emission in carbon monoxide observed in SN~1998S. We deduce a CO mass of $\\sim$ 10$^{-3}$ M$_{\\odot}$ moving at $\\sim$2200~km/s, and infer a mixed metal/He core of about 4~$M_{\\odot}$, again indicating a massive progenitor. Only three core-collapse supernovae have been observed in the $K$-band at post-100~days and all three have exhibited emission from CO. ", "introduction": "Core-collapse supernovae (SNe) are believed to arise from massive progenitors (M$\\geq$8-10M$_\\odot$). The lower-mass ($\\sim$ 8-10 M$_{\\odot}$) and higher-mass ($>20$ M$_{\\odot}$) progenitors experience heavy mass loss during the final stages of their evolution, several solar masses being ejected via a range of mass-loss rates. Consequently, in the vicinity of some core-collapse supernovae, dense circumstellar material (CSM) would be distributed according to the mass-loss history of the progenitor. The subsequent interaction of the freely-expanding supernova ejecta with the slowly-moving CSM generates a fast shock wave in the CSM and a reverse shock wave in the ejecta. The shocked regions emit high-energy radiation. The intensity of this emission depends on the density of the CSM and the ejecta, and the shock acceleration of the ejecta during the explosion (Chevalier \\& Fransson 1994). If the density of the CSM is small then the effects of the interaction only become significant several years after the explosion when the supernova has faded. However, if the CSM near the supernova is relatively dense, strong CSM-ejecta interaction can begin shortly after the explosion. This is the case for type IIn (n=narrow line) supernovae. These events exhibit narrow emission lines in their spectra superimposed on broader emission profiles. They also exhibit a strong blue continuum (Schlegel 1990). However, the broad P~Cygni absorption components typical of normal type~II SNe are weak or absent in type~IIn SNe. The presence of variable, narrow line emission is a direct manifestation of the excitation of the dense CSM by the SN radiation. In addition, the presence of broad H$\\alpha$ emission without a broad P Cygni absorption component is a clear indication that the observed broad line emission is powered by the ejecta-wind interaction (Chugai 1990). Consequently, type~IIn supernovae can provide unique information about the progenitor and its later evolution, through the observed properties of their CSM. In addition, the study of the interaction of these supernovae with the CSM can provide vital clues about galaxy evolution and the nature of active galactic nuclei ({\\it c.f.} Terlevich {\\it et al.} 1992). SN~1998S is the brightest type~IIn event ever observed. It was discovered on 1998 March 2.68 UT in the highly-inclined Sc galaxy NGC 3877 by Z. Wan (Li \\& Wan 1998) at a broadband (unfiltered) optical magnitude of +15.2. The supernova is located at 16$^{\\prime\\prime}$ west and 46$^{\\prime\\prime}$ south of the nucleus. By March 18.4 it had brightened to $V=+12.2$ (Fassia {\\it et al.} 2000). In a prediscovery frame obtained on 1998 February 23.7 there was no evidence of the supernova, down to a limiting apparent magnitude of $\\sim$+18 (Leonard {\\it et al.} 2000, IAUC 6835). We can thus assume that SN~1998S was discovered within a few days of the shock breakout. In the present paper we have adopted the discovery date 1998 March 2.68 UT = JD 2450875.2 as epoch 0 days, t$_{0}$, and express all other epochs relative to this fiducial date. Optical spectra obtained about day~3 (March~5--6) by Filippenko \\& Moran (1998), Huchra (Garnavich {\\it et al.} 1998) and ourselves (see $+$3.3 d spectrum in Figure~\\ref{figop}) showed prominent H and He emission lines with narrow peaks and broad wings superimposed on a blue continuum. As mentioned above these narrow lines indicate the presence of a dense CSM in the vicinity of the supernova. Using optical spectropolarimetry obtained at 5~days, Leonard {\\it et al.} (2000) [L00] deduced that the CSM is asymmetrically distributed. Bowen {\\it et al.} (2000) have presented high resolution spectroscopy of interstellar and circumstellar lines towards SN~1998S. They suggest that the CSM comprises a dense shell expanding at ~50~km/s with a more highly ionised shell moving at $\\sim$300~km/s. They also estimate that SN~1998S is at a real distance of $\\sim$10~kpc from the nucleus and deduce from the interstellar lines that it lies on the far side of the galaxy disk. Gerardy {\\it et al.} (2000) [G00] presented near-IR spectra of SN~1998S spanning 95--355~days post-maximum light. They identified emission from carbon monoxide. They also suggested that late-time multi-peak H and He line profiles in their optical and IR spectra indicate emission from a disk-shaped or ring-shaped circumstellar component. In addition, their t$\\geq$225~d near-infrared spectra exhibit a continuum that rises towards the longer wavelengths. They propose that the rising continuum is likely due to dust heated by the interaction of the ejecta with the CSM. In Fassia {\\it et al.} (2000) we presented contemporaneous optical and infrared photometric observations of SN~1998S covering the period between 11 and 146 days after discovery. Using the interstellar Na~I~D lines we derived an extinction $A_{V}=0.68^{+0.34}_{-0.25}$ mag. We also examined the evolution of the total luminosity and found that during the first month the luminosity decreased very rapidly {\\it viz.} $\\sim$0.08~mag/day. Subsequently (30-70 days) the decline rate decreased, resembling that of type~IIL supernovae {\\it viz.} 0.05 mag/day. By day~$\\sim$100 the decline rate slowed to 0.01 mag/day matching the radioactive decay of $^{56}$Co. From the bolometric luminosity after $\\sim$100 we estimate that 0.15$\\pm$0.05~M$_{\\odot}$ of $^{56}$Ni were produced in the explosion. Furthermore, we discovered that as early as day 130 the supernova exhibited an astonishingly high infrared (IR) excess, $K-L'=+2.5$. We argue that this excess is due to dust grains in the vicinity of the supernova. However, the physical extent of this early IR luminosity source was so large that the emission must have come from pre-existing dust in the CSM, possibly powered by X-rays from the ejecta-CSM interaction. In this paper we present optical and infrared spectroscopy of SN~1998S covering the period 3--127~days after discovery. The observations and the reduction procedure are presented in section~2. In section~3 we discuss possible line identifications. In section~4 we present an overview of the spectral behaviour of SN1998S and derive constraints about the nature and the characteristics of the CSM. We also analyse and model the first overtone of the CO emission. The work is summarised in section~5. ", "conclusions": "We have presented and discussed optical and infrared spectra of the type~IIn SN~1998S, covering epochs from a few days after explosion to over 100~days later. Our observational coverage of SN~1998S makes a significant contribution to the study of the type~IIn phenomenon. Of particular note is our acquisition of (a) contemporary spectra in both the optical and IR bands at a range of epochs, and (b) the most extensive set of high resolution spectra ever for this type of supernova event. The spectroscopic evolution of SN~1998S was complex. It can be understood in terms of the interaction of the supernova with a two-component progenitor wind which we have referred to as the ICSM and OCSM. Collision of the ejecta with the ICSM accounts for the early spectral features. From the fading of the broad emission components by $\\sim$14d we deduce that the outer boundary of this wind lay at less than 90~AU from the centre. Estimation of the inner wind velocity is difficult due to (a) the ongoing interaction of the shock with this wind even at the earliest of times, and (b) the lack of high resolution spectra at these times. We can only say that the early-time spectra exhibited features of width less than 400~km/s. However, these may have arisen in the OCSM. If, like L00, we assume that the inner and outer winds had similar velocities then from our high resolution measurements of the OCSM we can infer that the ICSM wind commenced less than 9~years ago. However, the possibility of an RSG-BSG evolution would argue against this assumption. Examination of the spectra indicates that the OCSM extended from 185~AU to over 1800~AU. Our high resolution spectra have revealed that the OCSM has a velocity of 40--50~km/s. Assuming a constant velocity during this time, we can infer that it commenced more than 170 years ago, and ceased about 20~years ago. During this period the outflow was high - at least $2 \\times 10^{-5}$M$_{\\odot}$~yr$^{-1}$, corresponding to a mass loss of at least 0.003~$M_{\\odot}$. An outflow of this strength and velocity is similar to those seen in cool supergiants. The broader absorption feature ($\\sim$350~km/s) in H and He~I may have arisen from a component of the outer CSM shell produced when the progenitor was going through a later blue supergiant phase. Alternatively, it may have been due to UV-precursor radiative acceleration of the inner part of the OCSM. Our analysis of the CO emission together with the bolometric light curve Fassia {\\it et al.} 2000) also indicates a massive progenitor, with a mixed metal/He core of $M\\sim4~M_{\\odot}$ i.e. comparable to that of SN~1987A. SN~1998S is only the third core-collapse SN for which post-100~day $K$-band spectra have been published, and yet all SN three events have exhibited first-overtone CO emission. (Fassia {\\it et al.} (1998) also published IR spectra for the core-collapse SN~1995V. No CO was identified but the latest $K$-band spectrum was only at 69~days.) A picture is therefore beginning to emerge where CO plays a ubiquitous role in the evolution of core-collapse SNe at late times. In particular, the powerful cooling property of CO is likely to lead to conditions in which dust may condense in the ejecta. Excess IR emission has been reported in SN~1998S on days~225, 260 and 255 (G00), day~253 (Garnavich {\\it et al.} 1998) and for a series of dates to as late as day~691 (Fassia et al. 2000), and this may be due to dust condensation. (An IR excess at 130~days was also reported by Fassia {\\it et al.} (2000) but they argue that it was unlikely to have arisen from dust condensation). In a future paper (Fassia {\\it et al.} in preparation) we shall examine in detail the evolution of the late-time IR excess. L00 used spectropolarimetry at early times to infer significant asymmetry in the OCSM and in the ejecta-ICSM interface. While our discussion of the high resolution OCSM lines assumed spherical symmetry, we have not provided quantitative explanations for the relative shifts between the allowed and forbidden lines. It may be that this will also require the introduction of an asymmetric model. G00 also deduced asymmetry in the OCSM, on the basis of the broad line profiles of H and He on days~240, 275, 370. We shall discuss these profiles in a future paper." }, "0011/astro-ph0011495_arXiv.txt": { "abstract": "We discuss the different definitions of the mass of a halo in common use and how one may convert between them. Using N-body simulations we show that mass estimates based on spherical averages are much more tightly correlated with each other than with masses based on the number of particles in a halo. The mass functions pertaining to some different mass definitions are estimated and compared to the `universal form' of Jenkins et al.~(\\cite{JFWCCEY}). Using a different simulation pipeline and a different cosmological model we show that the mass function is well fit by the Jenkins et al.~(\\cite{JFWCCEY}) fitting function, strengthening the claim to universality made by those authors. We show that care must be taken to match the definitions of mass when using large N-body simulations to bootstrap scaling relations from smaller hydrodynamical runs to avoid observationally significant bias in the predictions for abundances of objects. ", "introduction": "One of the most fundamental predictions of a theory of structure formation is the number density of objects of a given mass, the mass function, at a given redshift. Accurate mass functions are used in a number of areas in cosmology; in studies of galaxy formation, in measures of volumes (e.g.~galaxy lensing) and in attempts to infer the normalization of the power spectrum and the density parameter from the abundance of rich clusters. In the latter case the mass function is the point of contact allowing us to bootstrap the excellent statistics of large N-body simulations with observable properties of clusters normalized for example by hydrodynamic simulations of smaller volumes. In this way we can obtain reliable estimates of e.g.~the number of clusters as a function of temperature and redshift, which can in turn be used to constrain the matter density, the normalization of the power spectrum and the statistics of the initial density field. Several different definitions for the ``mass'' of a halo are in common use, each having different advantages. For example there are at least 3 different definitions of mass for the mass-temperature relation as computed from hydrodynamical simulations of galaxy clusters and all are different from the definition of mass commonly used in the mass function which is itself different from the mass usually employed in analytic studies based on the Press-Schechter~(\\cite{PS}; hereafter PS) theory. Observational data have improved to the point where it is important to distinguish between these different definitions of mass, lest we bias our theoretical predictions. We give an example of how such a conversion can be made, at least approximately, for a certain class of mass estimator. ", "conclusions": "We have shown that different definitions of the ``mass'' of a halo exist, and have different strengths and weaknesses. It is important to be consistent when combining relations which use different definitions of mass, and we have given an approximate method for converting between some commonly used mass estimators. Mass estimates based on spherical averages are much more tightly correlated with each other than with the mass obtained simply by summing the particles in the group, and can be quite well estimated by assuming a `universal' spherical profile (\\S\\ref{sec:corr}). The hydrodynamical simulations which calibrate observables as a function of cluster mass typically use such a spherically averaged mass definition. Unfortunately these definitions are not very tightly correlated with the particle based mass used in the universal mass function of halos reported by Jenkins et al.~(\\cite{JFWCCEY}). We have argued that this is because, with a linking length of $b=0.2$, FOF is merging neighbouring halos. Such a problem would be mitigated by reducing $b$, but it has not been demonstrated that the mass function is universal for $b\\ne 0.2$. In the cosmological model we have simulated, the fitting form of Jenkins et al.~(\\cite{JFWCCEY}) provides a good match to the mass function of our halos, strengthening the claim of those authors that it is universal. However we find the best fit when the mass estimator used is the top-hat virial mass, rather than the FOF mass. While we have not investigated other cosmological models, we expect that the profiles (from the point of view of estimating masses) will not be very cosmology dependent. It would be interesting to see whether the Jenkins et al.~(\\cite{JFWCCEY}) mass function is `universal' if one uses the (cosmology dependent) top-hat virial mass of the halos. \\bigskip I would like to thank V.~Springel for the use of his FOF group finder, J.~Mohr for useful conversations on this issue, and C.~Metzler for encouraging me to write it up. M.~White was supported by the US National Science Foundation and a Sloan Fellowship. Parts of this work were done on the Origin2000 system at the National Center for Supercomputing Applications, University of Illinois, Urbana-Champaign." }, "0011/astro-ph0011176_arXiv.txt": { "abstract": "\\vskip .2 truein Distance--redshift relations are given in terms of associated Legendre functions for partially filled beam observations in spatially flat Friedmann-Lema\\^\\i tre-Robertson-Walker (FLRW) cosmologies. These models are dynamically pressure-free, flat FLRW on large scales but, due to mass inhomogeneities, differ in their optical properties. The partially filled beam area-redshift equation is a Lame$^{\\prime}$ equation for arbitrary FLRW and is shown to simplify to the associated Legendre equation for the spatially flat, \\ie $\\OO=1$ case. We fit these new analytic Hubble curves to recent supernovae (SNe) data in an attempt to determine both the mass parameter $\\OM$ and the beam filling parameter $\\nu$. We find that current data are inadequate to limit $\\nu$. However, we are able to estimate what limits are possible when the number of observed SNe is increased by factor of 10 or 100, sample sizes achievable in the near future with the proposed SuperNova Acceleration Probe satellite. ", "introduction": "\\label{sec-intro} Distance-redshift or equivalently the Hubble curve is critical in determining current values of the cosmological parameters $H_0, \\OM$, and $\\OL$. Conversely, current values of these three parameters determine the large scale dynamics of the Universe into the distant past. A complication occurs when attempting to determine these parameters from high $z$ comparisons to the standard Hubble curve. The standard Hubble curve is a theoretical quantity computed assuming all gravitating matter is homogeneously distributed; whereas, observational data is taken in the real inhomogeneous Universe. In an inhomogeneous universe an observing light beam is lensed by inhomogeneities located external to, but near the light beam, and defocused (relative to the standard Hubble curve) by the less than average matter density within the beam. The simplest way to take into account these effects is to correct all beams for the missing homogeneous matter but correct for lensing only when necessary. This procedure requires the introduction of one additional parameter, \\eg a filling parameter $\\nu, \\ 0\\le\\nu\\le2$ defined by the fraction of inhomogeneous matter $\\rho_I/\\rho_0 \\equiv\\nu(\\nu+1)/6 \\le 1$ excluded from observing beams ($\\nu=0$ is the standard 100\\% filled beam FLRW case and $\\nu=2$ is the empty beam case). When observing high $z$ objects ($z\\sim 1$) the reader can think of the parameter $\\nu$ as representing matter that exists in galaxies but not in the intergalactic medium. To find the theoretical Hubble curve for observations in such a universe one must solve the geometrical optics equation [see \\cite{KR98}] given as equation (\\ref{Area}) in the next section. This equation is actually equivalent to the Lame$^{\\prime}$ equation for general FLRW but as pointed out by \\cite{KKT} reduces to the associated Legendre equation (\\ref{Legendre}) for the special case considered here, $\\OO=1$. In \\S\\,\\ref{sec-lumdist} we solve this equation using appropriate boundary conditions and give the Hubble curve in terms of associated Legendre functions (eq.[\\ref{Pans}]) as well as in terms of hypergeometric functions (eq.[\\ref{2F1ans}]). In \\S\\,\\ref{sec-fit} we fit this new Hubble curve to data for 60 supernovae (SNe) from the Supernova Cosmological Project (SCP) and from the Cala$^{\\prime}$n/Tololo Supernova Survey (CTSS) in an attempt to determine the mass parameter $\\OM$ and the filling parameter $\\nu$. In \\S\\,\\ref{sec-conclusions} we give some concluding remarks. ", "conclusions": "\\label{sec-conclusions} We have given useful forms for the luminosity distance in the currently relevant inhomogeneous $\\OO=1$ FLRW cosmologies. These cosmologies are all dynamically FLRW in the large but differ in how gravitating matter affects optical observations. A beam filling parameter $\\nu, \\ 0\\le\\nu\\le2$ allows the matter to vary from completely transparent and homogeneous to completely inhomogeneous and exterior to any observing beam. In order to determine the values of the cosmic parameters ($\\OM$, $\\OL$) from Hubble curves at high redshift, the value of $\\nu$ must also be constrained. For fixed values of ($\\OM, \\OL$), increasing $\\nu$ from 0 (totally homogeneous universe) to 2 (totally clumped) increases the distance moduli of points on the Hubble curve, {\\it especially at higher redshifts} as pointed out in \\S \\ref{sec-fit}. When observational error is taken into account, the problem of using standard candles at high redshift while ignoring $\\nu$ to obtain $\\OL$ will become particularly confounding. The current sample of SNe Ia at $z < 1$ fail to constrain the value of the beam-filling parameter $\\nu$. Samples 10 to 100 times larger than the current sample, and in the same redshift range, will constrain $\\nu$. In order to unambiguously determined ($\\OM$,$\\OL$) from even higher redshift observations like those planned in the future, the distance-enhancing effect of $\\nu$ must be accounted for in the luminosity distance formulae." }, "0011/astro-ph0011426_arXiv.txt": { "abstract": "We discuss the systematic effects arising from the {\\it cosmological redshift-space (geometric) distortion} on the statistical analysis of isodensity contour using high-redshift catalogs. Especially, we present a simple theoretical model for isodensity statistics in cosmological redshift-space, as a generalization of the work by Matsubara (1996). The statistical quantities considered here are the two- and three-dimensional genus of isodensity contour, the area of surface, the length of contour intersecting with a plane and the number of the crossing points of isodensity contour on a line. We give useful analytic formulae for the isodensity statistics, which take into account the corrections from the geometric distortion, the nonlinear clustering and the nonlinear velocity distortion phenomenologically. We then demonstrate how the geometric distortion and the nonlinear corrections alter shapes of the statistical quantities on the basis of plausible cosmological models. Our results show that the nonlinear correction can be sensitive to a choice of the redshift-space coordinate as increasing the redshift. The low-dimensional quantities such as two-dimensional genus systematically yield anisotropy due to the geometric and velocity distortions and their angle-dependence shows the $10\\sim20\\%$ difference of amplitude. Sensitivity for typical high-redshift samples are estimated in an analytic manner, and the influence of the light-cone effect for the isodensity statistics is also discussed. A simple estimation suggests that the systematic effects of geometric and redshift-space distortions can become comparable or could be dominant to the statistical error of deep cluster samples and future high-redshift galaxy surveys. These systematic effects might be a useful tool in probing the cosmological model of our universe. ", "introduction": "\\label{sec: intro} Recent observations of high-redshift objects have revealed dynamical aspects of the evolving universe. Measurements of Lyman-break galaxies provide evidences for the early galaxy assembly at the epoch $z\\sim 3$, which put a serious constraint on the structure formation scenario (Steidal et al. 1998; Giavalisco et al. 1998). Clusters up to $z\\sim 1$ have been observed, and the abundance suggests the universe with a low density parameter $\\Omega \\sim 0.3$ (e.g, Carlberg et al. 1997). Moreover, the upcoming wide-field surveys such as the Sloan Digital Sky Survey (SDSS) and the Two-Degree Field (2dF) survey promise to extend the observable scale of the universe and these enormous data will precisely reveal the clustering feature of the high-redshift quasars and galaxies. Since the analysis combining different redshift data sets has a potential to break the degeneracy of the cosmological parameters, as well as to provide the strong constraints on the structure formation scenario, statistical studies using the high-redshift objects are of great importance as a complementary to the measurements of cosmic microwave background anisotropy and the detection of cosmic shear. Recently several authors have discussed various cosmological effects arising from proper observations of the high-redshift objects (e.g, Suto et al. 1999 and references therein). Apart from the biasing of the luminous objects, the main effects are summarized as follows: \\begin{enumerate} \\item geometric distortion : no one can obtain the correct three-dimensional map of the high-redshift objects without knowing the correct cosmological model of the universe. The uncertainty of the cosmological model may affect proper evaluation of the clustering pattern in the {\\it cosmological redshift space} (Alcock \\& Paczy\\'nski 1979). \\item redshift-space distortion : peculiar motion of a cosmological object affects the estimation of the distance in a redshift survey, which causes the linear redshift-space distortion due to the bulk motion and the nonlinear velocity distortion due to the virialized random motion (finger-of-God effect) (Davis \\& Peebles 1983; Kaiser 1987; Suto \\& Suginohara 1991; Hamilton 1998 for a review). \\item light-cone effect : because an entire sample does not consist of objects on a constant-time hypersurface but on a light cone, the redshift evolution of the objects inevitably contaminates the data. Therefore, in a wide and deep survey, it becomes important to incorporate the light-cone effect into statistical quantities, for example, two-point correlation function and the power spectrum (Matarrese et al. 1997; Yamamoto \\& Suto 1999). \\end{enumerate} For the popular two-point statistics, the significance of these effects have been studied in detail and the robustness of the theoretical prediction is checked by the numerical simulation (Ballinger et al. 1996; Matsubara \\& Suto 1996; Magira et al. 2000; Hamana et al. 2000). Using the theoretical formulae, the feasibility of cosmological test with geometric distortion has been investigated (Yamamoto, Nishioka \\& Taruya 2000). Further, the analysis is developed to the higher-order moments on a light-cone (Matsubara et al. 1997). Here we extend these analyses to the isodensity statistics. That is, we investigate the influences of the above cosmological effects on the isodensity statistics, which have not been ever clarified. The isodensity statistics like the genus $G_3$ and area statistic $N_3$ have progressively become important as a quantitative measure of the topology of the large-scale structure. Analytical expressions of the isodensity statistics are derived for the Gaussian random field and some cases of non-Gaussian field in real space (e.g, Bardeen et al. 1986; Hamilton et al. 1986; Matsubara 1994, 2000; Matsubara \\& Yokoyama 1996). With these analytic formulae, the isodensity statistics computed from the redshift surveys of galaxies and/or the simulated dark matter catalogues are comprehensively understood (Colley et al. 2000 and references therein). While most authors of these works have focused on the low-redshift objects except for the cosmic microwave background anisotropy, the analysis using the high-redshift samples is expected to provide an important clue on the origin and evolution of proto-galaxies and proto-clusters. In fact, Park et al. (2000) have recently analyzed the two-dimensional genus using the galaxies taken from the Hubble Deep Field and quantified the feature of early galaxy assembly. Remarkably, their samples are widely distributed over the redshift range $0\\simlt z \\simlt 3$, which revealed an unprecedentedly deep view of galaxy clustering. Although the analysis is still limited to the projected density field because of the large error of photometric redshift, precise measurement of redshift will enable us to quantify the various isodensity statistics defined in the three-dimensional space, as well as the quantities in the one- and two-dimensional space in near future. Practically, the observed distribution of high-redshift objects suffers from the cosmological effects, which should be correctly taken into account when comparing such data with theoretical prediction. Then naive but important questions are as follows: how the cosmological effects affect the statistical study of isodensity contour; whether these effects are observed as signal or noise. Purpose of this paper is to address these issues by using a simple theoretical formulae for the isodensity statistics, in which the cosmological effects are faithfully incorporated. Partially, the influence of the linear (redshift-space) distortion is investigated by Matsubara (1996), though the effect turns out to be small. In his paper the corrections from the nonlinear clustering and the nonlinear velocity distortion, which become influential on small length scales, are not investigated, and neither is the geometric distortion. In this paper, we generalize the Matsubara's results to those incorporating the geometric distortion effect and the nonlinear correction taking account of the smoothing effect. With the extended formulae, we examine how the geometric and redshift-space distortions affect the isodensity statistics qualitatively and quantitatively. We will show that the nonlinear effect can be important even for a case with large smoothing scale because the scaling effect of the geometric distortion causes apparent enhancement of the nonlinear effects, which depends on choice of the cosmological redshift space. In addition we also discuss the sampling noise and the effect of finite redshift depth for typical high-redshift samples by using our extended formulae. We organized the paper as follows. In section \\ref{sec: Cosmological redshift space}, the cosmological redshift-space coordinate is introduced, and the statistical quantities incorporating the nonlinear corrections are described. Then we consider the isodensity statistics in cosmological redshift space in section \\ref{sec: isodensity}. Based on the analytic formulae for the Gaussian field in appendix \\ref{appendix: gaussian results}, we present a nonlinear model of isodensity statistics. After demonstrating predictions of the model, sensitivity of the isodensity statistics and the effect of the light-cone are discussed in section \\ref{sec: discussion}. The final section \\ref{sec: conclusion} is devoted to summary and conclusion. ", "conclusions": "\\label{sec: conclusion} In the present paper, we have considered various observational effects on the isodensity statistics of high-redshift objects, and estimated their influences, as a generalization of the work by Matsubara (1996). We have presented a simple theoretical model incorporating the geometric distortion and the nonlinear correction arising from the gravitational growth and the finger-of-God effects, together with useful analytic formulae. From this theoretical model (\\ref{eq: G_log})--(\\ref{eq: N1_log}), we have demonstrated that the geometric distortion affects the shapes of the isodensity statistics and that the nonlinear correction can be sensitive to a choice of the redshift-space coordinate as increasing the redshift. Based on these results, we have briefly discussed the detectability of these effects by evaluating the shot-noise and the sampling noise contaminations. Furthermore, the influence of the light-cone effect has been investigated. The important conclusions are summarized: \\begin{enumerate} \\item When the cosmological redshift space $s=\\shb(z)$ (eqs.[\\ref{eq: shb}] and [\\ref{eq: z_crd}]) is chosen, the nonlinear correction has an effect on the isodensity statistics for smoothing length larger $\\rs\\simeq 10h^{-1}$Mpc even at the higher redshift $z\\simgt1$, where the isodensity statistics have asymmetric shapes. On the other hand, when the cosmological redshift space $s=\\seds(z)$ (eqs.[\\ref{eq: seds}] and [\\ref{eq: EdS_crd}]) is chosen, the nonlinear correction turns out to be small and the linear theory of redshift-space distortions using the Gaussian results (\\ref{eq: G3_Gauss}), (\\ref{eq: G2_Gauss}) and (\\ref{eq: N3_Gauss})-(\\ref{eq: N1_Gauss}) could be a good approximation for the large smoothing scales. In both cases, the geometric and velocity distortions cause the angle-dependence in the low-dimensional quantities, and the relative amplitude can be changed by $10\\sim20\\%$. \\item The Poisson sampling error is crucial for the sparse sampling of high-redshift objects like the (SDSS/2dF) quasars. The finiteness of the sampling volume might affect the precise measurement of isodensity statistics with Lyman-break galaxies. Though more quantitative estimation of statistical errors using numerical simulations is needed, however, the isodensity statistics can be tested using clusters and high-z galaxies from future wide-field surveys. For these samples, the cosmological effects can be comparable or could be dominant to the Poisson and the sampling noise. \\item For a sample that the survey volume is large and the cosmological objects are distributed in a wide range of redshift, the light-cone effect can be influential. However, we infer that the light-cone effect has a weak influence on the isodensity statistics as long as the time evolution of the probability distribution function is not drastic, though this picture might depend on the evolution of the biasing factor. \\end{enumerate} Because the model presented here are quite simple, the predictions should be checked by numerical simulations. The prediction can also be sophisticated by using more reliable nonlinear models deduced from the simulations. Recent development of cosmological N-body simulation enables us to obtain the enormous large volume data over a few Gpc (Colberg et al. 2000). The larger data comparable to the Hubble volume size will provide us with chances of quantitative studies for the isodensity statistics, as well as the two-point statistics, in which the light-cone effect inevitably becomes important. Combining our theoretical template with these simulations, various cosmological implications might be yielded. One of the most interesting application is the geometric test proposed by Alcock \\& Paczy\\'nski (1979). As we have seen in Sec.\\ref{subsec: results}, the geometric distortion makes the observed structures distorted and it causes deformation of the isodensity contour. Since this effect is sensitive to the distance-redshift relation, the anisotropy of isodensity statistics might be a useful tool in investigating cosmological models, as well as the anisotropy of two-point statistics (Yamamoto, Nishioka \\& Taruya 2000). The present theoretical formulae include useful tools for quantifying the geometric distortion, as well as for testing statistical nature of biased density fields. To address these issues, however, a more quantitative analysis should be developed. The uncertainty of the biasing can be a critical issue to clarify the topological features of large-scale structure. Throughout the analyses, we have put the linear biasing ansatz (\\ref{eq: bias}) and the evolution of the biasing simply given by the model (\\ref{eq: FryBias}) neglecting the merging and formation processes. In more realistic situation, the biasing of luminous objects is nonlinear and stochastic (Dekel \\& Lahav 1999) and the merging and formation processes might alter the evolution of biasing (e.g, Somerville et al. 2000; Taruya et al. 2000). However, recent theoretical and numerical studies suggest that the linear biasing assumption (\\ref{eq: bias}) can be a good approximation even in the quasi-linear regime (Coles, 1993; Scherrer \\& Weinberg 1998; Matsubara 1999; Taruya \\& Suto 2000; Taruya et al. 2000). Therefore, in a qualitative sense, we believe that the present analyses would have revealed the generic feature of isodensity statistics on the observed high-redshift objects. As for the evolution of biasing, the theoretical prediction based on the Press-Schechter formalism gives the biasing factor for the dark halo distribution accurately (Mo \\& White 1996; Jing 1998; Sheth \\& Tormen 1999). Incorporating this evolution model into our analytic formulae can provide a powerful tool for quantifying the distribution of clusters of galaxies. In any way the isodensity statistics may probe a distinctive aspect of the biasing factor. This will be discussed elsewhere (Hikage, Taruya \\& Suto 2000, in preparation). \\bigskip \\bigskip We thank Prof. Suto for useful discussion and critical comments, Prof. Matsubara for useful discussion. A.T. gratefully acknowledges support from a JSPS (Japan Society for the Promotion of Science) fellowship. K.Y. thanks Prof. Kojima and Mr. Nishioka for useful comments. This work is supported by the Inamori foundation and in part by the Grants-in-Aid program (11640280) by the Ministry of Education, Science, Sports and Culture of Japan. \\clearpage" }, "0011/astro-ph0011083_arXiv.txt": { "abstract": "A numerical study of nonlinear $r$-modes in isentropic, rapidly rotating relativistic stars, via 3-D general-relativistic hydrodynamical evolutions, is presented. On dynamical timescales, we find no evidence for strong coupling of $r$-modes to other modes at amplitudes of order one or larger. Therefore, unless nonlinear saturation sets in on longer timescales, the maximum $r$-mode amplitude is of order one. An absolute upper limit on the amplitude is set by causality. Our simulations also show that $r$-modes and inertial modes in isentropic stars are discrete modes, with no evidence for the existence of a continuous part in the spectrum. ", "introduction": "The study of the properties of $r$-modes in rotating compact stars and their relevance to relativistic astrophysics has received considerable attention since the discovery (Andersson 1998, Friedman \\& Morsink 1998) that these modes are unstable to the emission of gravitational radiation. The $r-$mode instability provides an explanation for the spin-down of rapidly rotating young neutron stars to Crab-like spin-periods and for the spin-distribution of millisecond pulsars and accreting neutron stars. Additionally, it is considered to be a strong source of continuous gravitational radiation (see Friedman \\& Lockitch 1999 for a review). Moreover, if $r$-modes induce differential rotation, their interaction with the magnetic field in neutron stars can enhance the toroidal magnetic field of the star (Rezzolla, Lamb \\& Shapiro 2000). Before the instability can have an effect on the spin evolution of a young neutron star, the $r$-mode grows to an amplitude where it is saturated by nonlinear effects. Motivated by the absence of studies of such nonlinear saturation we performed a numerical study of $r$-mode hydrodynamical evolutions in rapidly rotating relativistic stars. We tried to elucidate what is the maximum amplitude such modes can reach, before nonlinear saturation (via hydrodynamical coupling) sets in (Stergioulas \\& Font 2000). The present contribution highlights the main findings of our study. We note that the saturation is most likely to set in on a hydrodynamical timescale, although it cannot be excluded that weak hydrodynamical couplings saturate the $r$-mode amplitude on longer timescales (but shorter than the growth timescale due to gravitational radiation reaction). However, at present, those long timescales cannot be achieved accurately in 3-D simulations, even with the largest available supercomputers. ", "conclusions": "\\begin{figure} \\centerline{ \\psfig{file=font_f1.eps,width=6.2cm,height=6.4cm} \\psfig{file=font_f2.eps,width=6.2cm,height=6.85cm} } \\caption{{\\it Left}: Evolution of the axial velocity in the equatorial plane for an amplitude of $\\alpha=1.0$, at $r=0.75R$. The evolution is a superposition of (mainly) the $l=m=2$ $r$-mode and several inertial modes. The amplitude of the oscillation decreases due to numerical (finite-differencing) viscosity of the code. {\\it Right}: Fourier transform of the velocity time-evolution, showing the frequencies of the modes in the inertial frame. The frequencies are the same at different radii, which implies a discrete spectrum.} \\label{fig_2} \\end{figure} Figure \\ref{fig_2} (left panel) displays the evolution of the axial velocity in the equatorial plane ($v^z$ along the $y$-axis) at a radius of $r\\sim 0.75R$. The evolution is a superposition of several modes, the $l=m=2$ $r$-mode being the dominant component. The chosen amplitude of the eigenfunction is $\\alpha=1.0$ (in the Newtonian limit, our definition of the amplitude agrees with that of Owen et al 1998). The perturbed star is evolved for more than 25ms (26 $l=m=2$ $r$-mode periods), during which the amplitude of the oscillation decreases due to numerical viscosity only. With much larger amplitudes the evolution is still similar to that in Figure \\ref{fig_2}, with no sign of nonlinear saturation of the $r$-mode amplitude on a dynamical timescale. Therefore, unless nonlinear saturation sets in at timescales much longer than the dynamical one, unstable $r$-modes could be driven to large amplitudes (of order one) by gravitational radiation reaction. An absolute upper limit on the $r$-mode amplitude is set by causality, requiring $\\sqrt{v_iv^i} 0.5$ mag. Our results indicate variations in $E_{V-I}$ of up to $\\sim$ 0.2 mag. Differential reddening this strong can wreak havoc with photometric and spectroscopic studies of cluster stars. The inherent uncertainties in the various parameters of the CMD (e.g., magnitude and color of the main-sequence turnoff (MSTO), horizontal branch features, etc.) are greatly amplified. Moreover, one may ``detect'' age or surface-temperature gradients where there are none in reality. To give an example of the magnitude of the effect, using the recently published color-temperature relations by Houdashelt, Bell, \\& Sweigart (2000),\\markcite{houdashelt00} the effective temperature of a solar-metallicity, main sequence star with $V-I \\sim 0.8$ would vary by approximately 600$K$ for a differential reddening effect of $E_{V-I} \\sim 0.2$ mag, which, in turn, could lead to errors in the metallicity determination. Using our internal dereddening technique, we obtain a high-quality, deep-photometry CMD of NGC 3201 comprised of approximately ninety 600s exposures as well as some shorter ones, all using the $V$ and $I$ bands. Our observations and data reductions, as well as the details about our internal dereddening method, are documented in Section 2. Section 3 contains our results concerning the reddening map and the CMD of NGC 3201. Finally, we discuss the determination of the reddening zeropoint of our extinction map in Section 4 and give a brief summary of our work in Section 5. ", "conclusions": "In the process of finding eclipsing binary stars in NGC 3201, we noticed the existence of variable reddening of up to 0.2 mag in $E_{V-I}$ on a scale of arcminutes. Using our internal dereddening method outlined in Section 2.2, we created an extinction map which is shown in Figures \\ref{extmap} and \\ref{extmapgrid}. Applying the map to our raw data (Fig. \\ref{cmd_raw}) significantly improved the appearance of the CMD (Fig. \\ref{cmd_final}). Comparison between our extinction map and the SFD map of the same region (Fig. \\ref{schlegelmap}) showed that the same larger-scale features exist in both maps. Our map displays some additional, smaller-scale features which are absent in the SFD maps (see Fig. \\ref{data-map}). The $E_{V-I}$ zeropoint which needs to be added to the numbers in Fig. \\ref{extmapgrid} in order to get absolute $E_{V-I}$ values is 0.15. This value is below literature results (Cacciari 1984, Harris 1996) by approximately $1.5 \\sigma$, but produced by far the best VDB isochrone fit to the data. The zeropoint determined with the help of the SFD maps gives $E_{V-I} \\sim 0.49$ as the average value across NGC 3201 which is higher than literature values which supports the statement by Arce \\& Goodman (1999) that the SFD maps overestimate the reddening in regions of high extinction. The results from this work will be essential in our binary star research where high-quality photometry of every binary system is necessary for distance determinations. A vital condition to obtaining these measurements is of course the knowledge of the exact extinction the star under investigation is suffering. Furthermore, studies like this may be useful in determining properties of the ISM itself such as examining a possible dependence of $R_V$ upon position in the field of view or giving insight into the distribution and properties of the dust along the line of sight." }, "0011/astro-ph0011574_arXiv.txt": { "abstract": "First results from high-resolution coronal spectroscopy of flares with the Reflection Grating Spectrometers on board the \\textit{XMM-Newton} satellite are reviewed. Rotational modulation in the X-ray light curve of HR 1099 is discussed. Results from time-dependent spectroscopy of flares in the active stars HR 1099, AB Dor, YY Gem are also presented. Variations in the shape of the emission measure distributions, in the abundances and in the average density of the cool plasma are discussed. ", "introduction": "Stellar coronae often display variability in their X-ray emission. Energetic explosive events (flares), eclipses, and rotational modulation are among the most interesting features found in X-ray light curves of coronal stars. Flares are at the center of a scenario that assumes an ensemble of flares as agents of heating of stellar coronae (``microflare hypothesis''; e.g., Parker~1988). Recent results suggest that flares can contribute significantly to the coronal heating of active stars (Audard et al.~2000). In a standard model, magnetic energy is released during a flare, heating the dense layers of chromospheric material. Hot ionized material is then driven into the corona by the pressure increase and fills the coronal loops (chromospheric evaporation; see Antonucci et al.~1994). Optically thin X-ray radiation occurs through continuum processes and electronic transitions. Flares are therefore important agents that bring ``fresh'' material from the chromospheric layers to the corona. In the non-flaring Sun, a First Ionization Potential (FIP) effect is observed, with enhanced coronal abundances for elements with low FIP ($<10$~eV), while high-FIP elements show solar photospheric values. In the stellar case, evidence for the presence of the FIP effect is less clear, but has been found in some stars (e.g., Drake et al.~1995; 1997). Recently, Brinkman et al.~(2001) has found in HR 1099 evidence for an \\emph{inverse} FIP effect and a strong enhancement of Ne relative to O. Stellar coronal flares often show enhanced metal abundances (e.g., Ottmann \\& Schmitt~1996). Also, some classes of solar flares can be Ca-rich (Ca is a low-FIP element; Sylwester et al.~1984) or Ne and S-rich (Ne and S are high-FIP elements; Schmelz 1993). In this paper, we report a summary of first results (see also G\\\"udel et al., this volume) from studies of stellar flares with the \\textit{XMM-Newton} Reflection Grating Spectrometers (RGS). Recent observations include HR 1099, AB Dor, YY Gem/Castor (see Audard, G\\\"udel, \\& Mewe~2001 and G\\\"udel et al.~2001ab for more details). ", "conclusions": "Flares are frequent in the stellar coronal sources observed so far by \\textit{XMM-Newton}. Large flares occurred on HR 1099, AB Dor, and YY Gem, allowing us to perform time-dependent spectroscopy with the high-resolution RGS in the range 5--35~\\AA\\ (0.35--2.5~keV). The presence of very hot (up to several tens of MK) material has been inferred, and evidence for elemental abundance enhancements has been found. Low-FIP elements appear to increase significantly during a flare on HR 1099, while the high-FIP element Ne stays at constant abundance. This behaviour contrasts with the \\emph{inverse} FIP effect found in these coronae during quiescence. Our \\textit{XMM-Newton} observations thus indicate that flares are also important as agents that relate chromospheric/photospheric plasma with coronal plasma. The flare EM distributions appear to be composed of a very hot plasma (up to $100$~MK) that evolves rapidly, and of a stable quiescent plasma. This result is consistent with the absence of a change in the density of the cool material. Stellar flares are important probes for stellar coronal plasma and the coronal heating mechanism. The very high temperatures attained by these flares are unknown to solar flares. Such observations are important to extend solar knowledge to more extreme conditions appropriate for magnetically active stars." }, "0011/astro-ph0011097_arXiv.txt": { "abstract": "We present a general linear dispersion relation which describes the coupled behavior of magnetorotational, photon bubble, and convective instabilities in weakly magnetized, differentially rotating accretion disks. We presume the accretion disks to be geometrically thin and supported vertically by radiation pressure. We fully incorporate the effects of a nonzero radiative diffusion length on the linear modes. In an equilibrium with purely vertical magnetic field, the vertical magnetorotational modes are completely unaffected by compressibility, stratification, and radiative diffusion. However, in the presence of azimuthal fields, which are expected in differentially rotating flows, the growth rate of all magnetorotational modes can be reduced substantially below the orbital frequency. This occurs if diffusion destroys radiation sound waves on the length scale of the instability, and the magnetic energy density of the azimuthal component exceeds the non-radiative thermal energy density. While sluggish in this case, the magnetorotational instability still persists and will still tap the free energy of the differential rotation. Photon bubble instabilities are generically present in radiation pressure dominated flows where diffusion is present. We show that their growth rates are limited to a maximum value which is reached at short wavelengths where the modes may be viewed as unstable slow magnetosonic waves. We also find that vertical radiation pressure destabilizes upward propagating fast waves, and that Alfv\\'en waves can be unstable. These instabilities typically have smaller growth rates than the photon bubble/slow modes. We discuss how all these modes behave in various regimes of interest, and speculate how they may affect the dynamics of real accretion disk flows. ", "introduction": "The physical state of the radiation pressure dominated, innermost regions of accretion disks around black holes has been uncertain ever since the early days of accretion disk theory. Standard alpha disk models in which the viscous stress is assumed to scale with the radiation pressure are subject to thermal and viscous instabilities (Lightman \\& Eardley 1974, Shakura \\& Sunyaev 1976). These instabilities are sensitive to the assumed prescription for the anomalous viscosity (e.g. Piran 1978), so it remains unclear whether or not they are actually present in real flows. If they are, the accretion flow may adopt a radically different state from that usually envisaged in thin accretion disk theory, such as the multiphase equilibrium recently proposed by Krolik (1998). In addition to these secular instabilities, dynamical instabilities also exist. First and foremost, a differentially rotating flow with a negative angular velocity gradient and an initially weak magnetic field is unstable to the magnetorotational instability (MRI, Balbus \\& Hawley 1991). The turbulence resulting from this instability is currently the most plausible mechanism known for generating the anomalous viscosity required in accretion disk models. Dissipation of magnetohydrodynamic waves excited by this turbulence by photon diffusion and photon viscosity has recently been examined by Agol \\& Krolik (1998). Photon diffusion might also affect the linear development of the instability itself in a laminar, radiation pressure dominated flow, but this issue has not been examined previously. Gammie (1998) has suggested that the overstable photon bubble modes discussed by Arons (1992) in the context of X-ray pulsars also exist in radiation dominated accretion flows in general. However, his instability analysis was limited to studying a static equilibrium where the effects of differential rotation were entirely neglected. Pietrini \\& Krolik (2000) have recently investigated convective instabilities in unmagnetized, radiation pressure dominated, differentially rotating flows. They assumed a constant vertical density profile, which necessarily leads to an unstable entropy gradient. Whether or not such unstable gradients exist in reality depends on the vertical dissipation profile, which in turn depends on the structure of the MHD turbulence. These dynamical instabilities may all play a role at some level in the dynamics and thermodynamics of the radiation pressure dominated portion of accretion disks. It is the purpose of this paper to provide a unified description of all three by deriving a general linear dispersion relation which incorporates them all. Such an analysis will hopefully prove to be a useful guide to numerical simulations which explore the nonlinear development of these instabilities and their effects on the resulting turbulent state of the inner accretion disk. This paper is organized as follows. In section 2 we discuss our basic equations and assumptions. Then in section 3 we focus exclusively on the MRI by deriving its dispersion relation in the absence of vertical stratification so that the other instabilities are suppressed. We are therefore able to study just the effects of photon diffusion on what is probably the most important of these instabilities. In section 4 we then present the full dispersion relation for a vertically stratified, differentially rotating medium, and discuss its solutions. We discuss the relevance of these solutions to astrophysical accretion disks in section 5 and then summarize our conclusions in section 6. ", "conclusions": "We may summarize our conclusions as follows. In the absence of an azimuthal field, the vertical ($k_R=0$) MRI modes have the standard, incompressible behavior elucidated by Balbus \\& Hawley (1991). If the azimuthal field component dominates the vertical field component and the Alfv\\'en speed is larger than the effective sound speed [cf. eq. (\\ref{vaphicond})], then the growth rate of the MRI is reduced to $\\sim\\Omega(\\cgas/\\vaphi)$, well below the usual rapid growth rate of order the orbital frequency $\\Omega$. However, the instability still persists. Large vertical photon fluxes in the equilibrium do not significantly affect the most rapidly growing MRI modes. At short wavelengths where rotation is unimportant, both the fast and slow MHD waves are destabilized by sufficiently high vertical photon flux in the equilibrium. These instabilities require that the waves propagate at some nonzero angle to the vertical, and affects upward propagating fast waves and downward propagating slow waves. The slow waves appear to always have the larger growth rates. At smaller wavenumbers these are well-described by Gammie's (1998) photon bubble dispersion relation (\\ref{pbdisp}), but at larger wavenumbers they asymptote to a maximum value given by equation (\\ref{slowdisp}). At small wavenumbers where rotation is important ($k1600$ d may form through accreting part of the ejecta from the intrinsic AGB stars through stellar wind, and the mass accretion rate is in the range of 0.1-0.5 times of Bondi-Hoyle's accretion rate. Those with shorter orbital period $P<600$ d may be formed through dynamically stable late case C mass transfer or common envelope ejection. ", "introduction": "Since Burbidge et al. (1957) and Cameron (1957) published their creative studies, the nucleosynthesis theory has been developed in deep degree. In particular, asymptotic giant branch (hereafter AGB) stars are very important to study element nucleosynthesis and the Galactic chemical evolution because they synthesize significant parts of slow process (hereafter s-process) neutron capture elements and $^{12}$C. The products are taken out from the stellar interior, He-intershell, to the surface by the third dredge-up (hereafter TDU) process, and then are ejected into interstellar medium with the progressive stellar wind mass loss. Our understanding of the AGB nucleosynthesis has undergone major revisions in these years. The earlier studies (Iben 1975; Truran \\& Iben 1977) illustrated that intermediate mass TP-AGB stars with $^{22}$Ne$(\\alpha,n)^{25}$Mg neutron source (the typical neutron density is least of a order of $10^{9}-10^{10} n$ cm$^{-3}$) are the suitable nucleosynthesis sites of s-process elements. But new observations shed doubts on the above idea (Busso et al. 1995 and references therein; Busso et al. 1999 and references therein). Busso et al. (1995) and Lambert et al. (1995) demonstrated that the measured abundances of Rb/Sr, the products of the branch in the s-process path at $^{85}$Kr, imply a definitely lower neutron density (typical of the order of $10^{7}n$ cm$^{-3}$), which can be provided via the reaction $^{13}$C$(\\alpha,n)^{16}$O at low temperature in the He-intershell of low mass AGB stars. Iben \\& Renzini (1982a,b) indicated that a suitable mechanism operated in low mass stars of low metallicity to allow the formation of a semiconvective layer, hence the $^{13}$C pocket. The pocket is engulfed by the next convective pulse where $^{13}$C nuclei easily capture $\\alpha$ nuclei, release neutrons. Hollowell \\& Iben (1988) confirmed the possibility of formation of a consistent $^{13}$C pocket through a local time-dependent treatment of semiconvection. However, the semiconvection mixing mechanism was not found to work for the $^{12}$C-enriched Population I red giants, like the peculiar stars of MS, S and C stars, showing overabundances of s-process elements in their spectra. Straniero et al. (1995) investigated the effect of a possible mixing of protons into a thin zone at the top of the carbon-rich region during each dredge-up episode, hence the formation of $^{13}$C pocket. They suggested that the $^{13}$C was completely burnt in the radiative condition, and the resulting s-process nucleosynthesis occurs during the quiescent interpulse period, instead of the convective thermal pulse. $^{22}$Ne$(\\alpha,n)^{25}$Mg was still active for a very short period during the convective pulse with minor influence on the whole nucleosynthesis. Herwig et al. (1997) and Herwig et al. (1998) supported the formation of $^{13}$C pocket via hydrodynamical calculations. Straniero et al. (1997) adopted the above new s-process nucleosynthesis scenario to calculate the s-process nucleosynthesis of solar metallicity low mass AGB stars with $1 \\leq M/M_{\\odot} \\leq 3$, and gave the detailed results. Recently, Gallino et al. (1998) explained further and developed the aforesaid new scenario. They divided the $^{13}$C pocket, $q$ layer, into three zones in the light of the distribution in the mass of hydrogen introduced in the $^{12}$C-rich intershell. The characteristic neutron exposures in the three layers are different. Moreover, when the nucleosynthesis occurs in a radiative layer, only the nucleosynthesis products are ingested into the convective thermal pulse, which makes the classical concept of mean neutron exposure ($\\tau_0$) become meaningless and the simple assumption of an exponential distribution of the neutron exposure fail to account for the complexity of the phenomenon (Arlandini et al. 1995, Gallino et al. 1998). Busso et al. (1999) reviewed this new s-process nucleosynthesis scenario in details. The spectral and luminosity studies of AGB stars (including MS, S and N-type C stars) have shown that the M$\\rightarrow$S$\\rightarrow$C sequence is the result that the low-mass AGB stars have undergone carbon synthesis, s-process nucleosynthesis and the third dredge-ups (Lambert 1991). These stars are in the course of experiencing thermal pulse stage, and the original chemical abundances of their atmospheric envelope have been modified by two mixing mechanisms, namely, the first dredge-up when they became red giants and the third dredge-up when they became TP-AGB stars (Boothroyd $\\&$ Sackmann 1988a, b, c and d). The predicted evolutionary sequence of M$\\rightarrow$S$\\rightarrow$C in the heavier$-$lighter s-element abundances relationship (here and hereafter, 'heavier' refers to the second metal peak elements: Ba, La, Ce, Nd and Sm etc.; 'lighter' refers to the first metal peak elements: Y, Zr etc.) and the heavy-element abundances$-$C/O relationship ($^{12}$C, together with the s-process elements, is dredged up from stellar interior during the third dredge-up) are important to understand the nucleosynthesis and evolution of AGB stars. Because (1) they will provide theoretical basis for the observed evolution of the sequence, (2) they can check the available theories on the evolutions of AGB stars (e.g., the beginning of the third dredge-up process, the mass and the chemical abundance of the dredge-up material, the theory of s-process nucleosynthesis, the stellar wind mass loss, and the formation of carbon stars etc.). Busso et al. (1992) discussed the heavy-element abundances of M, MS and S stars using the thermal pulse AGB model. Busso et al. (1995) analyzed the heavy-element overabundances of carbon stars under the assumption that the dredge-up started after reaching the asymptotic distribution (about the 20th pulse). It is difficult to calculate the AGB stars evolution and s-process nucleosynthesis. So there are few theoretical results to explain the M$\\rightarrow$S$\\rightarrow$C evolutionary sequence based on the combination of the heavier$-$lighter s-element abundances ratio and the C/O ratio, though the observational abundances exhibit a certain regularity. Zhang et al. (1998a) calculated the evolution of the surface heavy element abundances and C/O ratio for a 3$M_{\\odot}$ TP-AGB star with initial metallicity 0.015, and gave interesting results. But they adopted mean neutron exposure $\\tau_0$ and unbranch s-process path, which have been revised in these years. In the first part of this paper, we adopt the new s-process nucleosynthesis scenario (Straniero et al. 1995; Straniero et al. 1997; Gallino et al. 1998; Busso et al. 1999 etc.), and the branch s-process nucleosynthesis path to calculate the s-process nucleosynthesis of solar metallicity 3$M_{\\odot}$ AGB stars. And then, we discuss the M$\\rightarrow$S$\\rightarrow$C sequence based on the heavy-element abundances and C/O ratio. The importance of AGB stars nucleosynthesis is not only to explain the observational M$\\rightarrow$S$\\rightarrow$C sequence but also to be responsible for the origin of some other classes of stars with overabundances of heavy-elements. Observations revealed that some stars with overabundances of heavy-elements were not luminous to up to the stage of AGB. Following Lambert (1991), the stars showing heavy-element overabundances are divided into two classes: (1) intrinsic TP-AGB stars$-$they include MS, S and C (N-type) stars exhibiting the unstable nucleus $~^{99}$Tc ($\\tau_{1\\over 2}=2\\times 10^5 ~$yr) as evidence that they are presently undergoing nucleosynthesis activity and the third dredge-up, and (2) extrinsic AGB stars$-$they include the various classes of G-, K-type barium stars and the cooler S, C stars where $~^{99}$Tc is not observed. It is generally believed that the extrinsic AGB stars belong to binary systems and their heavy-element overabundances come from accretion of the matter ejected by the companions (the former AGB stars, now evolved into white dwarfs) (McClure et al. 1980; Boffin \\& Jorissen 1988; Jorissen et al. 1998; Jorissen \\& Van Eck 2000; Jorissen 1999). The mass exchange took place about $1\\times 10^6$ years ago, so the $^{99}$Tc produced in the original TP-AGB stars have decayed. The accretion may either be disk accretion (Iben $\\&$ Tutukov 1985) or common envelope ejection (Paczynski 1976). Han et al. (1995) detailedly investigated the evolutionary channels for the formation of barium stars. In this paper we will only discuss the stellar wind accretion scenario because it is very important to explain the formation of barium stars (Boffin \\& Jorissen 1988; Jorissen et al. 1998). Boffin $\\&$ Jorissen (1988) calculated qualitatively the variation of orbital elements caused by wind accretion in binary systems. They also estimated the heavy-element overabundances of barium stars. Subsequent papers (Za$\\check{c}$s 1994, Boffin $\\&$ Za$\\check{c}$s 1994) used similar methods to calculate the overabundances, and interpreted the relationship between the heavy-element abundances and the orbital periods of barium stars. Some important conclusions have been drawn in the theory of wind accretion, but the previous calculations on orbital elements were not very reliable because of the neglect of the $\\delta r/r$ term ($r$ represents the distance between the two components of the binary system ), and using the tangential momentum conservation (Chang et al. 1997 and references therein). For the rotating binary system with wind mass loss, the total angular momentum conservation is more reasonable than tangential momentum conservation. Also, the previous calculations of heavy-element overabundances used the 'step-process' (Boffin $\\&$ Jorissen 1988, Boffin $\\&$ Za$\\check{c}$s 1994), which means that the overabundance factor changes at one single instant from 1 to $f$ ($f$ is the relative ratio of the heavy-elements to the solar abundances, and differs for different elements. Earlier calculations used the mean $f$ value of carbon stars), and then keeps the value until the end of the AGB phase. However, after the start of the TDU, the overabundance factor $f$ of the intrinsic AGB stars changes during successive dredge-ups. It is after a number of dredge-ups that the C/O ratio in the outer envelope of the intrinsic AGB star reaches the value 1, which means that the AGB star becomes a carbon star. The heavy-element overabundances of the barium star should be caused by the successive pulsed accretions and mixing. According to the analysis to the orbital elements of barium and S stars, Jorissen \\& Mayor (1992) presented the evolutionary pathways of binaries leading to barium and S systems. They concluded that the binary systems with longer orbital period formed through wind accretion and those with shorter orbital period formed via Roche lobe overflow. But the specific range of orbital period was not presented. Jorissen et al. (1998) analyzed the orbital elements of a large sample of binary systems to give insight into the formations of barium and Tc-poor S stars. They suggested that barium stars with orbital period $P$$>$1500 days formed through wind accretion scenario. Zhang et al. (1999) suggested that the barium stars with $P$$>$1600 days formed via wind accretion according to their model. Liu et al. (2000) discuss the evolutions of orbital elements of barium stars from normal G, K giants. Besides the orbital elements, the heavy-element abundances of barium stars have been discussed in some literatures. Busso et al. (1995) discussed the observational heavy-element abundances of barium stars. And a more detailed analysis of the abundance distributions for five stars has been performed using the method of mixing the accretion mass with the envelope mass. But the effect of mass accretion and the changes of orbital elements were not considered. Chang et al. (1997) and Liang et al. (1999) attempted to explore the relationships between the heavy-element overabundances and orbital elements of bariums stars using the binary accretion scenario, but with the shortcomings: or using the tangential momentum conservation, or adopting the old nucleosynthesis scenario of TP-AGB stars. In the second part of this paper, we firstly reduced the variation equations of orbital elements based on the angular momentum conservation model of wind accretion, then we calculate the heavy-element overabundances of barium stars via successive pulsed accreting matter enriched heavy-elements from the intrinsic AGB stars, and mixing the matter with their envelopes. This paper is organized as follows. The observational data of MS, S, C (N-type) stars and barium stars are given in Sect. 2. In Sect. 3, we present the model and the main parameters of AGB stars nucleosynthesis and the angular momentum conservation model of wind accretion scenario for barium stars. Sect. 4 illustrates and analyzes our results. We conclude and discuss in Sect. 5. ", "conclusions": "\\subsection * {5.1 The intrinsic AGB stars} Adopting the latest AGB stars evolutionary theory and nucleosynthesis scenario, the heavy-element abundances of solar metallicity 3$M_{\\odot}$ AGB stars are calculated. It is shown that, adopting reasonable parameters, the calculated results of the heavier/lighter s-elements ratio$-$ overabundance relationships and the C/O ratio$-$overabundance relationships can fit the observations. The evolution of AGB stars along the M$\\rightarrow $S$\\rightarrow$C sequence is thus further explained from the evolutionary theories and heavy-elements nucleosynthesis of AGB stars. A 3$M_{\\odot}$ AGB star will undergo about 27 pulses before it becomes carbon star. The results showed that the overabundances of s-process elements depend significantly on the neutron exposure: for the lower neutron exposures ($a$$<$1.5), the lighter s-process elements are more abundant than the heavier ([hs/ls]$<$0.0); for the higher neutron exposures ($a$$>$1.5), the latter can increase strongly with the increase of neutron exposures, and the lighter s-process elements change slightly ([hs/ls]$>$0.0); while the results with $a$=1.5 can explain the solar heavy-element abundance distribution, which confirm the results of Gallino et al. (1998). Moreover, $^{12}$C abundance correlates to the s-process element abundances. Both C/O ratio and s-element abundances increase with the occurrence of TDU. After some dredge-ups, C/O ratio reaches 1, from which the AGB stars become carbon stars. The observed MS, S and C (N-type) stars lie in a range of neutron exposure: $a$=0.5-2.5. The over low neutron exposure can not produce carbon stars (e.g. $a$=0.5). Our calculation thus provides a further theoretical basis for the evolution of AGB stars along the M$\\rightarrow $S$\\rightarrow$C sequence based on the heavy-element abundances and C/O ratio. The general agreements between our calculations and the observations indicate that the parameters, the theories of evolution and nucleosynthesis scenario of AGB stars adopted in calculation are reasonable. The stars with initial main sequence mass 1.5$-$4$M_{\\odot}$ can form carbon stars (Groenewegen et al. 1995). We consider only a 3$M_{\\odot}$ star model in our calculation. So the comparison between the calculated results and the observations is somewhat rough. With further studies of the evolutionary theory of AGB stars and increasing observational data, we can expect a deeper understanding on the nucleosynthesis of AGB stars. \\subsection * {5.2 The barium stars} Taking the conservation of angular momentum in place of the conservation of tangential momentum for wind accretion scenario, considering the change of $\\delta r/r$ term, and not neglecting the square and higher power terms of eccentricity, the change equations of orbital elements are recalculated. We combine wind accretion with the nucleosynthesis of intrinsic AGB stars, to calculate, in a self-consistent manner, the heavy-element overabundances of barium stars through mass accretion during successive pulsed ejection, followed by mixing. The calculated relationships of heavy-element abundances $-$ orbital period $P$ can fit the observations within the error ranges. Moreover, the predictions of the detailed abundances of different atomic charge Z can match well the observations of 11 program barium stars with longer orbital period ($P$$>$1600 d). The corresponding neutron exposures are in the range of 0.8$<$$a$$<$2.5. The higher neutron exposure (e.g. $a$=2.0) will produce the more abundant the heavier s-elements than the lighter (see Fig. 4c); on the contrary, the abundances of the lighter s-elements are higher than the heavier with low neutron exposure (e.g. $a$=1.0, see Fig. 4a); while $a$=1.5 will produce the almost equal abundances of the heavier to the lighter s-elements (see Fig. 4h). These neutron exposures can fit the corresponding results of intrinsic AGB stars. Naturally, we can understand the observations of no-Tc MS and S stars, which are commonly believed to be the descendants of barium stars. These results of the extrinsic AGB stars confirm the reliability of the nucleosynthesis and evolution of intrinsic AGB stars. Simultaneously, the results confirm that our wind accretion model and parameters adopted are suitable. Analyzing our results, we understand that the barium stars with longer orbital period ($P$$>$1600 d) form through wind accretion. Those with shorter orbital period ($P$$<$600 d) form through other scenarios, such as dynamically stable late case C mass transfer or common envelope ejection. Moreover, the change range of mass accretion rate should be 0.1 to 0.5 times as much as the Bondi-Hoyle's accretion rate. The corresponding range of orbital periods and mass accretion rate to the formation of barium stars still need to be tested by more observations. At present, the orbital elements of a large sample barium stars, Tc-poor S stars have been published (Udry et al. 1998a, 1998b; Carquillat et al. 1998). But the corresponding heavy-element abundances have not been obtained. So we need the high resolution, high signal-to-noise spectral observations of these stars, which are combined with the observations of orbital elements, to research their characters and formation. In addition, we should note that metallicity is an important factor to the AGB stars nucleosynthesis and the formation of chemical peculiar stars. Actually, $^{13}$C neutron source is related to the metallicity (Busso et al. 1995, 1999 and references therein; Gallino et al. 1999). Gallino et al. (1998) calculated the s-element nucleosynthesis of 2$M_{\\odot}$ AGB stars with low metallicity $Z$=0.01, and obtained similar abundance distribution to the 3$M_{\\odot}$ with solar metallicity AGB stars model. Busso et al. (1999) and Zhang et al. (1998b) have discussed the inverse correlation between the heavy-element abundances and the metallicity [Fe/H] of intrinsic and extrinsic AGB stars. Also, the nucleosynthesis results of low metallicity AGB stars are more suitable to study the Galactic chemical evolution in the early stage of the Galaxy. For extrinsic AGB stars, Jorissen et al. (1998) suggested that the s-process operation was more efficient in a low-metallicity population, so the Pop. II CH stars may has accreted the material much enriched in heavy elements from the former AGB companion. In this paper, our main aims are (1) calculating the AGB stars nucleosynthesis, so that we can explain the observed heavy-element abundances of MS, S and C (N-type) stars, which are near solar metallicity, and supply evidence to the M$\\rightarrow $S$\\rightarrow$C evolutionary sequence; (2) discussing the wind accretion scenario of Ba stars, which are near solar metallicity too (Za$\\check{c}$s 1994; Smith et al. 1993). So we only calculate the solar metallicity case. We will extend to study the low metallicity case in the forthcoming paper." }, "0011/astro-ph0011448_arXiv.txt": { "abstract": "I present a series of diagrams which illustrate why the cosmic microwave background (CMB) data favor certain values for the cosmological parameters. Various methods to extract these parameters from CMB and non-CMB observations are forming an ever-tightening network of interlocking constraints. I review the increasingly precise constraints in the $\\Omega_{m} - \\Omega_{\\Lambda}$ plane and show why more cosmologists now prefer $\\Lambda$CDM cosmologies to any other leading model. ", "introduction": "A convenient way to interpret CMB observations is to fit the angular power spectrum of the data to parameter-dependent models. Figure \\ref{fig:data} shows the recent CMB measurements along with three such models. In Figure~\\ref{fig:binnedata}, binning of this data reduces the scatter and provides a representative region favored by the data. Important parameters that can be constrained by CMB power spectra include Hubble's constant $h$, the cosmological constant $\\Omega_{\\Lambda}$, the density of cold dark matter $\\Omega_{\\rm CDM}$, and the density of baryonic matter $\\Omega_{b}$. Figures~\\ref{fig:data}~-~\\ref{fig:zoomin} provide a qualitative feel for the lever arm that the CMB data provides for constraining these and other parameters simultaneously. Unless stated otherwise, the models shown have the following default values: $h= 0.70$, $\\Omega_{\\Lambda}=0.7$, $\\Omega_{m}= \\Omega_{\\rm CDM} + \\Omega_{b} =0.3$, $\\Omega_{b}h^{2} = 0.020$, a power spectral index of primordial scalar density fluctuations $n_{s}=1$ and an overall normalization $Q_{10} = 18\\;\\mu$K. The grey band in Figure~\\ref{fig:binnedata} is reproduced in Figures~\\ref{fig:highh}~-~\\ref{fig:zoomin} and represents the data in a model-independent way. With it, the eye can pick out which models best fit the data. A reduction in $h$ increases the amplitude of the first peak (Fig.~\\ref{fig:highh}). An increase in the number of baryons increases the gravitating mass of the oscillating baryon-photon fluid . This enhances the first peak (Fig.~\\ref{fig:omegabaryon}) by producing more gravitational compression as the baryons drag the photons further into the potential wells (and further away from the potential hills). The second peak is suppressed because, before decoupling, these smaller scales experienced the same additional compression (and rarefaction) and, at decoupling, we are seeing a subdued rebound from this enhanced compression (and rarefaction), i.e., we are seeing the smaller amplitude of an oscillation whose zero level had been lowered in the previous oscillation by the effect of additional baryons. An increase in $\\Omega_{m}$ decreases the amplitude of the first peak (Fig.~\\ref{fig:breakdegeneracy}). For discussion of the physics of the parameter dependencies of features in the CMB power spectrum see e.g. Hu \\& Sugiyama (1995), Hu (1995), Tegmark (1996), Lineweaver et al. (1997). \\clearpage \\begin{figure}[!ht] \\plotfiddle{cl1.ps}{7.0cm}{0}{63}{53}{-187}{-57} \\caption{{\\footnotesize Compilation of recent CMB observations from 22 experiments compared to three popular cosmological models. At the last IAU, all three models were serious contenders for reality. This is no longer the case.}} \\label{fig:data} \\end{figure} \\begin{figure}[!hb] \\plotfiddle{cl1fill.ps}{7.0cm}{0}{63}{53}{-187}{-57} \\caption{{\\footnotesize Same as Fig.~\\ref{fig:data}, but the data points have been binned and averaged and rebinned multiple times to provide a series of points and a grey area defined by the 1 $\\sigma$ error bars of the binned points. The points are not independent. The well defined position of the first peak in the spectrum at $\\ell_{p} \\sim 220$ easily rules out the $(\\Omega_{m}, \\Omega_{\\Lambda}) = (0.3,0.0)$ model shown which peaks at $\\ell_{p} \\sim 400$ (see e.g. Lineweaver 1998).}} \\label{fig:binnedata} \\end{figure} \\clearpage \\begin{figure}[!ht] \\plotfiddle{cl1h.ps}{6.3cm}{0}{60}{48}{-175}{-52} \\caption{{\\footnotesize {\\bf Standard CDM is consistent with CMB data, but only if $h$ is extremely low.} Until recently, the standard cold dark matter model ($\\Omega_{m}, \\Omega_{\\Lambda}) = (1,0)$ was the leading cosmological candidate, but the amplitude of the first peak is too low in these models unless $h \\sim 0.40$. See the $h$ contours in Panel A of Figure~\\ref{fig:science} for more details. The grey region is preferred by the data and comes from the previous figure.}} \\label{fig:highh} \\end{figure} \\begin{figure}[!hb] \\plotfiddle{cl1obh2.ps}{6.3cm}{0}{60}{48}{-175}{-52} % \\caption{{\\footnotesize {\\bf Why does the CMB prefer high values of $\\Omega_{b} h^{2}$?} Compared to big-bang-nucleosynthesis-dependent estimates ($\\Omega_{b} h^{2} \\approx 0.020$ e.g. Tytler et al. 2001) the first peak of the CMB data has a slightly higher amplitude and the second peak has a slightly lower amplitude. Raising $\\Omega_{b} h^{2}$ to 0.030 fits the data better because it raises the first peak and lowers the second peak. More precise data in the $\\ell \\sim 500$ region will soon be available to help solve this tentative conflict. }} \\label{fig:omegabaryon} \\end{figure} \\clearpage \\begin{figure}[!ht] \\plotfiddle{cl1flath455565.ps}{6.5cm}{0}{60}{48}{-187}{-57} % \\caption{{\\footnotesize {\\bf Degeneracy is a big problem.} These three flat models ($\\Omega_{tot} = \\Omega_{m} + \\Omega_{\\Lambda} = 1, \\Omega_{k} = 0$) have very different values for $\\Omega_{m}$ and $\\Omega_{\\Lambda}$, but the CMB cannot differentiate between them if $h$ can vary . Reducing $h$ raises the peak (Fig.~\\ref{fig:highh}) while increasing $\\Omega_{m}$ lowers the peak. Similar but usually more subtle degeneracies become more numerous as the dimensionality of explored parameter space increases. An important issue is what range of values does one allow for $h$ and other parameters.}} \\label{fig:degeneracy} \\end{figure} \\begin{figure}[!hb] \\plotfiddle{cl1flath70.ps}{6.5cm}{0}{60}{48}{-187}{-57} % \\caption{{\\footnotesize {\\bf Degeneracies can be broken.} Precise non-CMB constraints (such as an independent measurement of $h$) or much more precise CMB data can break degeneracies. Here, setting $h = 0.70$ breaks the degeneracy of Fig. 5, allowing the CMB to discriminate between various flat models.}} \\label{fig:breakdegeneracy} \\end{figure} \\clearpage \\begin{figure}[!ht] \\plotfiddle{cl1omegakv2.ps}{5.8cm}{0}{57}{45}{-175}{-52} \\caption{{\\footnotesize {\\bf The CMB can measure the geometry of the Universe, $\\Omega_{k}$, better than other cosmological observations.} The position of the first peak in the CMB data, $\\ell_{p}$, is probably the best measurement we have of the geometry of the universe. It is an excellent tool to answer the question: `Is the universe spatially open, flat or closed?' Or, in terms of cosmological parameters ($\\Omega_{k} = 1- \\Omega_{tot}$), `Is $\\Omega_{k}$ greater than, equal to or less than zero?' (or equivalently, `Is $\\Omega_{tot}$ less than, equal to or greater than one?') We have set $\\Omega_{k} = 0.0, 0.2, 0.4$ (flat, open, more open, respectively). The vertical lines indicate the peak in the data (Fig.~\\ref{fig:zoomin}).}} \\label{fig:geometry} \\end{figure} \\begin{figure}[!hb] \\plotfiddle{cl1lpeakv4.ps}{5.8cm}{0}{57}{45}{-175}{-52} \\caption{{\\footnotesize {\\bf Where is the peak?} This is a blow-up of the region of the first peak with the multiply-binned points from Fig.~\\ref{fig:binnedata} and the models from Fig.~\\ref{fig:omegabaryon}. An eye-ball model-independent estimate of the position of the first peak is $\\ell_{p} \\approx 225 \\pm 25$, indicated by the vertical lines (same as in previous figure). This result is based on the data shown in Fig.~\\ref{fig:data}. The points are {\\it not} independent.}} \\label{fig:zoomin} \\end{figure} \\clearpage Just as the SNIa data is our strongest lever arm for determining the acceleration of the universe, $q_{o}$, the CMB data is our strongest lever arm for determining the geometry of the universe, $\\Omega_{k}$. The ability of the CMB to constrain $\\Omega_{k}$ can be seen in Fig.~\\ref{fig:science}, panel A and Fig.~\\ref{fig:latest} in which the CMB contours are elongated in the $\\Omega_{k} = $ {\\it constant} direction (upper left to lower right). However, since both $\\Omega_{k}$ and $h$ control the position of the peak, there is a slight degeneracy. It is difficult to separate the effect of the spatial geometry of the universe from the effect of $h$. This degeneracy is reflected in the width of the elongated CMB constraints in the $\\Omega_{m} - \\Omega_{\\Lambda}$ plane. The models in Fig.~\\ref{fig:zoomin} have $\\Omega_{k} = 0$, $h = 0.70$. The data and the high baryon model peak at $\\ell \\approx 225$. This crude eye-ball estimate should be compared to the more careful but model-dependent estimates of Bond et al. (2001, $\\ell_{p} = 212 \\pm 7$) and Hu et al. (2001, $\\ell_{p} < 218$ at the $2\\sigma$ level based on the Boomerang and Maxima results only). ", "conclusions": "" }, "0011/astro-ph0011162_arXiv.txt": { "abstract": "Hipparcos proper motions and trigonometric parallaxes allow the derivation of secular parallaxes which fix the distances to individual stars in the Hyades cluster to an accuracy of $\\sim$2~percent. The resulting color-absolute magnitude diagram for 92 high-fidelity single members of the cluster displays a very narrow main sequence, with two turn-offs and associated gaps. These occur at the locations where the onset of surface convection affects the $B-V$ colors, as predicted by B\\\"ohm--Vitense thirty years ago. The new distances provide stringent constraints on the transformations of colors and absolute magnitudes to effective temperatures and luminosities, and on models of stellar interiors. ", "introduction": "The Hyades open cluster is one of the key calibrators of the absolute magnitude-spectral type relation and the mass-luminosity relation. The small distance to the cluster ($\\sim$45~pc) has a number of advantages for studies of its Hertzsprung--Russell diagram: (i) the foreground interstellar reddening and extinction is negligible ($E(B-V) = 0.003 \\pm 0.002$~mag \\citep[e.g.,][]{tay1980}), (ii) the large proper motion ($\\mu \\sim 111$~mas~yr$^{-1}$) and peculiar space motion ($\\sim$35~km~s$^{-1}$) facilitate proper motion- and radial velocity-based membership determinations, and (iii) the stellar content can be probed to low masses relatively easily. Among the $\\sim$400 known members are white dwarfs, red giants, mid-A stars in the turn-off region, and main-sequence stars down to at least $\\sim$$0.10~M_\\odot$ M dwarfs \\citep*{bhj1994}. At the same time, the proximity of the Hyades complicates astrophysical interpretation: the tidal radius of $\\sim$10~pc corresponds to a significant depth along the line of sight. As a result, the precise definition and location of the main sequence and turn-off region in the Hertzsprung--Russell diagram, and the accuracy of the determination of, e.g., the helium content and age of the cluster, have always been limited by the measurement error of the distances to individual stars. Even Hipparcos parallax uncertainties translate into absolute magnitude errors of $\\ga$0.10~mag at the mean distance of the cluster, whereas $V$-band photometric errors only account for $\\la$0.01~mag uncertainty for most members. \\begin{figure*}[t] \\begin{center} \\includegraphics[angle=0.0, width=18.0truecm, clip=true, keepaspectratio=true] {debruijne01.eps} \\end{center} \\caption[]{Color-absolute magnitude diagram for 92 high-fidelity members of the Hyades cluster. This sample excludes all members beyond 40~pc from the cluster center, double and multiple stars, and stars with suspect secular parallaxes (see BHZ for details). The absolute magnitudes were computed using the observed $V$-band magnitudes and secular parallaxes derived by \\citet{bhz2000}. The $B-V$ colors were taken from the Hipparcos Catalog. The arrows indicate the two B\\\"ohm--Vitense turn-offs and associated gaps, which are most likely caused by sudden changes in the properties of convective atmospheres. The gaps between $B-V=0.15$ and $0.20$, between $B-V=0.30$ and $0.35$, and the gap around $B-V=0.95$ mag are caused by the suppression of double, multiple, and peculiar stars from our sample (cf.\\ figure 21 in \\citet{per1998}; see \\S 2); the region between $B-V=0.30$ and $B-V=0.35$ mag, e.g., is occupied by Am-type stars, which have a high incidence of duplicity \\citep[e.g.,][]{jj1987}.} \\label{fig:camd} \\end{figure*} It was realized long ago \\citep[e.g.,][]{bos1908} that the small internal velocity dispersion of the Hyades makes it possible to improve the trigonometric parallaxes of individual member stars by kinematic modelling of their proper motions. Various recent studies derived such secular parallaxes for Hyades members from Hipparcos astrometry \\citep*{per1998,bru1999,dlm1999,ng1999}, but none investigated the resulting Hertzsprung--Russell diagram in any detail. Yet, as the relative proper motion accuracy is effectively three times larger than the relative trigonometric parallax accuracy, the secular parallaxes are three times more precise than the trigonometric parallaxes, and provide a unique opportunity to obtain a well-defined and absolutely calibrated Hertzsprung--Russell diagram for an open cluster. For this reason we have redetermined the secular parallaxes for the Hyades, using the comprehensive procedure described by \\citet{bru1999} \\citep*[cf., e.g.,][]{lmd2000}. We describe the full derivation of the secular parallaxes elsewhere \\citep*[BHZ]{bhz2000}. This includes a new analysis of the space motion and internal velocity dispersion of the Hyades, a validation of the secular parallaxes, with a detailed investigation of the effect of velocity structure in the cluster and of the presence of small-angular-scale correlations in the Hipparcos data, and an extensive analysis of the construction of the physical Hertzsprung--Russell diagram $(\\log T_{\\rm eff}, \\log L)$. In the course of this work, we constructed a color-absolute magnitude diagram for a sample of high-fidelity single members of the Hyades, and discovered that the improved accuracy reveals two turn-offs and associated gaps in the main sequence at the locations predicted by B\\\"ohm--Vitense, in the seventies. These are the topic of this Letter. \\begin{figure*}[t] \\begin{center} \\includegraphics[angle=-90.0, width=8.0truecm, clip=true, keepaspectratio=true] {debruijne02.eps} \\end{center} \\caption[]{Schematic color-absolute magnitude diagram of an open cluster with a B\\\"ohm-Vitense gap. Stars left of the dashed vertical line, at $B-V\\sim 0.3$ mag, have radiative envelopes. Surface convection sets in at the line, and all stars to the right of it have convective envelopes. While the onset of convection does not alter luminosity or effective temperature, it causes redder colours because of lower temperatures in deeper layers (see \\S 3). The resulting color-absolute magnitude diagram therefore shows a turn-off and gap.} \\label{fig:bvgap} \\end{figure*} ", "conclusions": "Narrow main sequences in color-magnitude diagrams are readily observable for distant clusters, but the absolute calibration of the Hertzsprung--Russell diagram of such groups is often uncertain due to their poorly determined distances and the effects of interstellar reddening and extinction. The latter problems are alleviated significantly for nearby clusters, but at the price of introducing a considerable spread in the location of individual members in the Hertzsprung--Russell diagram as a result of their resolved intrinsic depths, and relatively poorly determined individual distances. We have shown here that the Hyades are unique in that the secular parallaxes derived from the Hipparcos astrometry for the members are sufficiently accurate to calibrate the main sequence of this nearest open cluster in absolute terms. The color-absolute magnitude diagram for the Hyades shown in Figure~\\ref{fig:camd} reveals two turn-offs and associated gaps in the main sequence. We identify these with the so-called B\\\"ohm--Vitense gaps, which are most likely related to sudden changes in the properties of surface convection zones in the atmospheres of stars with $B-V \\sim 0.30$ and $\\sim$0.45~mag, which significantly affect the emergent UV and blue-optical fluxes, and thus the $B-V$ color. We show in BHZ that this substructure in the $(B-V, M_V)$ diagram provides a strong constraint on stellar models, requiring an improvement in the available transformations from colors and absolute magnitudes to effective temperatures and luminosities. The future astrometric satellites FAME and GAIA will improve the accuracy of stellar astrometry into the micro-arcsecond regime. They will make it possible to provide accurate membership and high-precision absolutely-calibrated main sequences for all star clusters and associations to distances of at least 2 kpc. This will provide the ability to test stellar models over a range of metallicities and ages, and may well reveal further substructure." }, "0011/astro-ph0011481_arXiv.txt": { "abstract": "We present a method, based on the kinematic analysis of the Galactic disk stars, to clarify whether the internal motions of the stellar system in spiral arms follow those expected in the density wave theory. The method relies on the comparison with the linear relation between the phases of spatial positions and epicyclic motions of stars, as drawn from the theory. The application of the method to the 78 Galactic Cepheids near the Sun, for which accurate proper motions are available from the {\\it Hipparcos} Catalogue, has revealed that these Cepheids hold no correlation between both phases, thereby implying that their motions are in contradiction with the theoretical predictions. Possible reasons for this discrepancy are discussed and future prospects are outlined. ", "introduction": "Spiral structures of galaxies have been studied for a long time in order to understand how these structures are formed (e.g., Roberts, Roberts \\& Shu 1975; Rohlfs 1977; Binney \\& Tremaine 1987). One of the proposed models to explain spiral arms is that they are just material arms, where the stars originally making up a spiral arm remains in the arm even at the later time. However this simple model holds a wellknown problem which is called ``winding problem'': the differential rotation in galactic disks winds up the arm in a short time compared with the age of galaxies, so that the spiral pattern would be too tightly wound compared with the observed spiral structures. In contrast, the currently most popular model, which is free from the winding problem, is the density wave theory (Lin \\& Shu 1964), where a spiral arm is regarded as a wave and wavelike oscillation of stellar motions propagates through galactic disks. In this picture, the global spiral pattern is sustained independently of individual stars moving at different angular velocities. For a comprehensive review of the density wave theory, see, e.g., Rohlfs (1977) and Binney \\& Tremaine (1987). The density wave theory has been suggested by various observational aspects in spiral galaxies, including the relative distributions of dust lane, interstellar gas, and H II regions across the arms (Fujimoto 1968; Roberts 1969; Rohlfs 1977), intensity distribution of radio continuum radiation (Mathewson, van der Kruit, \\& Brouw 1972), and systematic variation of gaseous velocity fields near the arms (e.g., Visser 1980). In particular, recent high-resolution observations using CO emission have revealed detailed streaming motions of molecular gas, which are generally in agreement with predictions of the density wave models (e.g., Kuno \\& Nakai 1997; Aalto et al. 1999). However, we note that these observational results provide only an outcome of nonlinear interaction between interstellar matter and background stellar arms, and it is yet unknown whether the motions of stars {\\it themselves}, which make up spiral pattern, actually follow those predicted by the density wave theory. In this regard, the direct access to detailed stellar kinematics in disks is possible only in our Galaxy. Here, we present a method to clarify this issue, based on the analysis of local kinematics of disk stars. We then apply the method to 78 Cepheids in the solar neighborhood, for which the precise data of proper motions are available from the {\\it Hipparcos} Catalogue (ESA 1997). Also, the distances to these sample stars can be accurately estimated from the period-luminosity relation, so that combined with the radial velocity data, the full three-dimensional velocities are available. We note here that although we focus on the local kinematics of spiral arms in this work, the method we develop here can be applied to the motions of more remote stars distributed over a whole disk, for which precise astrometric data will be provided by the next-generation satellites such as {\\it FAME} and {\\it GAIA}. Our paper is organized as follows. In \\S 2, we describe the method to determine whether or not the motions of stars agree with those expected in the density wave theory. In \\S 3, we show the detail of the sample stars and the fundamental parameters of our Galaxy adopted in this work. The application of our method to the sample stars is shown in \\S 4. Finally, \\S 5 is devoted to discussion and conclusions. ", "conclusions": "We have presented a method, based on the analysis of local kinematics of disk stars, to clarify whether the motions of the stars, which make up the spiral arms, follow those expected in the density wave theory. The method utilizes the comparison with the expected linear relation between the ``position phases'' of the stars $\\chi$ and those of their epicyclic motions $\\phi$, as given in equation (\\ref{2}). The application of the method to the 78 Galactic Cepheids within 4 kpc from the Sun, for which accurate proper motions are available from {\\it Hipparcos}, has revealed that the relation between $\\chi$ and $\\phi$ for the Cepheids does not show the expected linear relation. Based on the quantitative analysis using the deviation $\\Delta^2$, we conclude that the observed motions of the Cepheids are well reproduced by the random model having no correlation between $\\chi$ and $\\phi$. There are a couple of possibilities to explain the current results, even if the spiral arms follow the density wave theory. First, the spiral structure around the Sun may not be simple as given in equation (\\ref{1}). In fact, many of our sample Cepheids belong to the local spiral structure called the Orion arm, for which the definite conclusion on its spatial structure is yet to be reached. It is frequently expressed as ``Orion spur'', having a rather irregular pattern compared to other large-scale arms, Sagittarius and Perseus arm (see, e.g., Gilmore, King, \\& van der Kruit 1989). The existence of the Orion arm may give disturbances on the density wave motions of stars induced by these large-scale arms. Second, the Cepheids we have adopted here may still convey systematic velocities of dens gas clouds from which these stars were formed, in the form of the streaming motions. If there still exist some individual streaming motions among the sample stars, such motions may violate the ideal linear relation between $\\chi$ and $\\phi$ expected for the density wave motions. Third, the Cepheids we have adopted here may have already experienced some scattering by dens gas clouds, thus having large velocity dispersions (Spitzer \\& Schwarzschild 1953). However, our experiment in \\S 4 (dotted line in Figure 4) implies that the effect of velocity dispersions of our sample on the result appears to be minor. In order to settle the last issue described above more clearly, we have repeated our analysis using younger populations with smaller velocity dispersions than the Cepheids. As such young stars, we have adopted the O-B5 stars, although due to their fainter luminosities than Cepheids, the sample with available proper motions is confined to the narrower region near the Sun. These sample stars are taken from the NASA SKY2000 Master Star Catalog Ver.~2 (Sande et al. 1998) which provides almost 300,000 stars brighter than 8 mag. The catalog contains many basic quantities, such as MK classification, luminosity class, apparent magnitude, color, radial velocity, and so on. We have then calibrated distances using {\\it Hipparcos} parallaxes or spectroscopic distances using the program kindly supplied by Drs. M.~S\\^oma and M.~Yoshizawa, and also obtained accurate proper motions by the cross-identification with the {\\it Hipparcos} and ACT Reference Catalogs (Urban, Corbin, \\& Wycoff 1998). After removing binaries and multiples, we have selected 773 O-B5 stars for which full three-dimensional velocities are available. Then, the application of the method we have developed here has revealed that the deviation $\\Delta^2$ as a function of $m$ remains essentially constant of the order of 3000, without showing any noticeable minimum at a specific value of $m$. Thus, even the motions of such young populations with small velocity dispersions are in contradiction with the density wave theory. We note here that most of these O-B5 stars are located within $\\sim 1$ kpc from the Sun, so the effect of the local irregular spiral on the result cannot be negligible. More definite conclusions on the issue we have addressed here require the assembly and analysis of much larger numbers of stars with accurate distances and proper motions, so that the statistical fluctuation in the result can be significantly reduced. Also, it is necessary to assemble the data of more remote stars over a large fraction of the disk, thereby diminishing effects of local irregular spiral structures on the kinematic analysis. Indeed, next-generation satellites such as {\\it FAME} and {\\it GAIA} will provide very precise astrometric data for huge numbers of the Galactic stars, and will thus offer us an opportunity to assess detailed motions of disk stars in conjunction with the density wave theory." }, "0011/astro-ph0011354_arXiv.txt": { "abstract": "Oscillation frequencies are the most accurate properties one can measure for a star, potentially allowing detailed tests of stellar models and evolution theories. We briefly review asteroseismology for two classes of stars. In $\\delta$~Scuti variables, the main problems are the identification of the observed modes and the theoretical treatment of rotation. In solar-like stars the main difficulty is the tiny amplitudes, but credible detections are now being made. These confirm that stars are oscillating at the approximately the expected levels, but suggest that amplitudes scale as $1/g$ rather than $L/M$. We also stress the importance of multi-site campaigns. Several space missions will be launched over the coming years, promising an exciting future for asteroseismology. ", "introduction": "Asteroseismology involves an interplay between observations of stellar oscillations and theoretical model calculations. It can be done when a set of oscillation frequencies is known for a given star and, at the same time, a set of theoretical model frequencies can be calculated. The motivation for doing asteroseismology is that oscillation frequencies are the most accurate properties one can measure for a star and we may therefore, at least in principle, be able to perform a detailed test of stellar modelling and evolution theories. This potentially very promising tool has motivated a huge observational effort with the aim of obtaining accurate oscillation frequencies. At the same time, a substantial amount of work has been put into improvements of stellar modelling with the goal of being able to fit model frequencies to the observed oscillation frequencies. In helioseismology, as can be seen from the papers presented in these proceedings, most current work concentrates on the details. We are already able to measure accurately many important properties of the solar interior. Unlike asteroseismology, helioseismology is founded on an immense amount of high-quality data supported by an equally detailed set of state-of-the-art models. It is impossible to imagine that asteroseismology will ever reach a level similar to that in which we find helioseismology today, concerning analysis techniques, data quality and the level of the results. The reasons for this lie in the main differences between the two subjects: \\begin{itemize}\\itemsep=0pt \\vspace*{-2ex} \\item Asteroseismology works on distant stars whose basic properties will always be less well determined than those of the Sun. These include age, composition, radius, mass, atmospheric properties and neutrino flux. The uncertainties will affect the quality of the calculated stellar models. One example is age, which is known for the Sun from radioactive dating of the solar system. To illustrate the importance of this parameter for the current solar model, consider the following simple question: What age would we determine for a solar model -- based on the observed properties of the Sun -- if we did not already know the age of the solar system? \\item The surfaces of stars are essentially unresolved, which limits asteroseismology to modes of low degree. Much of what we have learned about the Sun is based on high degree modes. \\item The Sun is a relatively simple system. This makes it very interesting, since we may have a chance to understand it! Many stars seem to be much more complicated. Hopefully we will find some important physical properties that are not known at present, which may turn out to be important for understanding the details of those stars. \\item There is only one Sun, while there are billions of other stars. Many people work on the details in helioseismology, but we can't expect several billion astronomers to be working in asteroseismology! \\item In helioseismology we see some big networks and dedicated telescopes (GONG, IRIS, BiSON), and we have many years of space research (SOHO, SolarMAX, IPHIR). In asteroseismology we find many smaller campaigns and networks, but very few dedicated telescopes. The space projects are just beginning. So far, there is a tremendous lack of high-quality data. \\item Asteroseismology works on fainter stars and so has lower sensitivity than helioseismic observations. One will therefore generally be restricted to oscillations with relatively high amplitudes. \\end{itemize} To explain why asteroseismology is so far behind helioseismology, we could identify three important elements from the above list: \\begin{itemize}\\itemsep=0pt \\vspace*{-2ex} \\item The imprecision of basic stellar properties. \\item The complication of the stellar physics. \\item The lack of high-quality oscillation data. \\end{itemize} ", "conclusions": "In this review we have described the current status of asteroseismology for A, F and G subgiants and main sequence stars. The main conclusions are as follows: \\begin{itemize} \\item Asteroseismology is far from helioseismology in terms of techniques and results. \\item There are plenty of challenges for theoreticians, such as mixing, fluid motions, turbulent convection and deviations from spherical symmetry. The most important area to focus on is rotation. \\item Asteroseismology is observationally driven. Ground-based velocity observations are now achieving believable detections, and it is time for multi-site campaigns. A large ground-based network is very desirable. \\item A number of exciting space missions are in various stages of planning and construction (MOST, COROT, MONS and Eddington). If these succeed, it seems likely that asteroseismology will enter a golden age. \\end{itemize}" }, "0011/astro-ph0011024_arXiv.txt": { "abstract": "We present new optical and ultraviolet spectroscopy of the anomalous EUV emitter HD~199143 (F8V). High resolution spectra in the H$\\alpha$ and Na\\,{\\sc i}\\,D wavelength regions show evidence for very rapid (a few hundred km~s$^{-1}$) rotation of the stellar photosphere. Using archive {\\it IRAS} data we also show that the star has excess emission above photospheric levels at 12~$\\mu$m. {\\it IUE} data of HD~199143 reveal the presence of emission lines of Mg\\,{\\sc ii}, C\\,{\\sc i}, C\\,{\\sc ii}, C\\,{\\sc iii}, C\\,{\\sc iv}, Si\\,{\\sc iv}, He\\,{\\sc ii} and N\\,{\\sc v} and show a large variability, both in the continuum and line fluxes. We propose that all available data of HD~199143 can be explained by assuming that is has been spun up by accretion of material from a close T~Tauri like companion, responsible for the emission lines, the ultraviolet variability and the excess infrared emission. The bursting or flaring nature of this object, mostly in high energies, could be explained as episodic mass transfer between the star and its close companion. We show that HD~199143 and the Li-rich late-type dwarf BD$-$17\\degr6128 form a physical pair and suggest that both may be part of a new nearby (48~pc) young ($\\sim$ 10$^7$ yr) stellar association in Capricornius. ", "introduction": "Zuckerman \\& Webb (2000) sketch a picture of the recent star formation history of the solar neighbourhood in which 10--40 million years ago an ensemble of molecular clouds were forming stars at a modest rate near the present position of the Sun. About 10~Myrs ago, the most massive of these newly formed stars exploded as a supernova, terminating the star formation episode and generating the very low density region seen in most directions from the present Sun (Welsh et al. 1998). This scenario can not only explain the presence of young stellar groups close to the earth, but also explains how the $\\beta$~Pic moving group can be so young (20~Myr; Barrado y Navascu\\'es et al. 1999), and yet so close. However, currently this scenario is largely speculative. HD~199143 is a poorly studied bright ($V$ = 7\\fm27) star in the constellation of Capricornius. It has been classified as F8V in the Michigan Spectral Survey (Houk \\& Smith-Moore 1988), after an initial classification of G0 by Cannon \\& Mayall (1949). The star would be completely inconspicuous, if it hadn't been detected as a bright extreme-ultraviolet source by the {\\it ROSAT} (2RE J205547$-$170622) and {\\it Extreme Ultraviolet Explorer} (2EUVE J2055$-$17.1) missions (Pounds et al. 1993; Malina et al. 1994; Pye et al. 1995; Bowyer et al. 1996). In this {\\it Letter} we present new optical and ultraviolet spectroscopy of HD~199143 and show that it is a variable and rapidly rotating F8V star. We infer that all characteristics of the HD~199143 system can be explained by assuming that it is a binary system, in which the primary has been spun up by accretion of mass from a low-mass companion. Its association with a previously studied T~Tauri-like system (BD$-$17\\degr6128) suggests that these two stars could be the first two members of a close (48~pc) new region of recent star formation and may provide compelling support for the star formation history of the solar neighbourhood outlined in the first paragraph. ", "conclusions": "The presence of a normal late-type companion could not explain the ultraviolet excess, or the {\\it EUVE} and {\\it ROSAT} detections of HD~199143. However, the presence of an accretion disk around our hypothetical companion, such as that found in LMXB or T~Tauri systems, might easily explain those properties, as well as the infrared excess, the presence of emission lines and the variability. In such a scenario, the high rotational velocity of HD~199143 could be due to a spin-up in its past by accretion from the companion. At first glance, a scenario in which a nearby main-sequence star like HD~199143 would have a T~Tauri-like companion would seem far-fetched. However, Mathioudakis et al. (1995) report the presence a strongly flaring K7e--M0e dwarf with a high Li abundance only 5 arcminutes from HD~199143. The optical spectrum of this star, BD$-$17\\degr6128, is identical to that of many T~Tauri stars. From Digital Sky Survey images we identify BD$-$17\\degr6128 with HD~358623. An inspection of the Tycho-2 Catalogue (H{\\o}g et al. 2000) shows that this star has a proper motion of 59 $\\pm$ 3 and $-$63 $\\pm$ 3 mas~yr$^{-1}$ in $\\mu_\\alpha$ and $\\mu_\\delta$, identical to that of HD~199143. From the fact that HD~358623 is the only star within a 5 degree radius for which this is the case, we exclude the possibility that this could be a coincidence and conclude that the two stars form a genuine proper motion pair. Using the data by Mathioudakis et al. (1995), and the newly determined distance, we compute the absolute luminosity of BD$-$17\\degr6128 to be 0.34 $\\pm$ 0.06 L$_\\odot$, employing a similar procedure to that followed for HD~199143. Comparison with the pre-main sequence evolutionary tracks by D'Antona \\& Mazzitelli (1997) yields an age of 10$^7$ years for BD$-$17\\degr6128, consistent with a T~Tauri nature of this star. Using the radial velocity of HD~199143 determined in Section~2, and the parallax and proper motions listed in the {\\it Hipparcos} catalogue, we compute the galactic space velocity components $(U, V, W)$ of HD~199143 to be $(-10 \\pm 13, -13 \\pm 6, -13 \\pm 6)$ km~s$^{-1}$. This space motion is similar to that of many stars in the vicinity of the Tucanae and TW Hydra associations (Zuckerman \\& Webb 2000), suggesting that these stars might have formed from the same cloud complex. We conclude that HD~199143 and BD$-$17\\degr6128 could very well be the first two members of a region of recent star formation similar to the TW Hydrae Association and the newly identified Tucanae Association (Kastner et al. 1997; Zuckerman \\& Webb 2000). If confirmed, a further study of these two enigmatic stars could lead to a better understanding of the star formation history in the solar neighbourhood." }, "0011/astro-ph0011212_arXiv.txt": { "abstract": "We create mock pencil-beam redshift surveys from very large cosmological $N$-body simulations of two Cold Dark Matter cosmogonies, an Einstein-de Sitter model ($\\tau$CDM) and a flat model with $\\Omega_0 =0.3$ and a cosmological constant ($\\Lambda$CDM). We use these to assess the significance of the apparent periodicity discovered by Broadhurst \\etal (1990). Simulation particles are tagged as `galaxies' so as to reproduce observed present-day correlations. They are then identified along the past light-cones of hypothetical observers to create mock catalogues with the geometry and the distance distribution of the Broadhurst \\etal data. We produce 1936 (2625) quasi-independent catalogues from our $\\tau$CDM ($\\Lambda$CDM) simulation. A couple of large clumps in a catalogue can produce a high peak at low wavenumbers in the corresponding one-dimensional power spectrum, without any apparent large-scale periodicity in the original redshift histogram. Although the simulated redshift histograms frequently display regularly spaced clumps, the spacing of these clumps varies between catalogues and there is no `preferred' period over our many realisations. We find only a 0.72 (0.49) per cent chance that the highest peak in the power spectrum of a $\\tau$CDM ($\\Lambda$CDM) catalogue has a peak-to-noise ratio higher than that in the Broadhurst \\etal data. None of the simulated catalogues with such high peaks shows coherently spaced clumps with a significance as high as that of the real data. We conclude that in CDM universes, the regularity on a scale of $\\sim 130h^{-1}$Mpc observed by Broadhurst \\etal has {\\it a priori} probability well below $10^{-3}$. ", "introduction": "The redshift distribution of galaxies in the pencil-beam survey of Broadhurst et al. (1990, hereafter BEKS) displayed a striking periodicity on a scale of 128$h^{-1}$Mpc. This result has attracted a good deal of interest over the subsequent decade, and the significance and nature of periodicity or regularity in the distribution of galaxies has remained the subject of a stimulating debate in both observational and theoretical cosmology. Although a number of studies have been devoted to the BEKS pencil-beam survey and other similar surveys, several fundamental questions remain unanswered. From the theoretical viewpoint, it is important to decide whether such apparently periodic galaxy distributions can occur with reasonable probability in a Cold Dark Matter universe, or require physics beyond the standard paradigm. Performing large simulations can directly address this question. The first simulation specifically designed for pencil-beam comparisons was that of Park and Gott (1991, hereafter PG). Their rod-shaped CDM simulation allowed them to create twelve quasi-independent mock pencil-beam surveys similar in length to that of BEKS. One of their samples appeared `more periodic' than the BEKS data according to the particular statistical test they used for comparison. Other authors (Kurki-Suonio et al.; Pierre 1990; Coles 1990; van de Weygaert 1991; SubbaRao \\& Szalay 1992) have used purely geometrical models such as cubic lattices and Voronoi foams to explore the implications of apparent regularities similar to those found by BEKS. In particular, SubbaRao and Szalay (1992) presented a sequence of Monte Carlo simulations of surveys of Voronoi foams, showing that such a model can successfully reproduce the data as judged by a variety of statistical measures, for example, the heights, positions and signal-to-noise ratios of the highest peaks in the power spectra. Kaiser \\& Peacock (1991) argued that the highest such peaks in the BEKS data are not sufficiently significant to be unexpected in a CDM universe, but did not support this conclusion with detailed simulations. Dekel et al. (1992) introduced other statistics, more similar to those of PG, and again concluded that the apparent periodicity seen in the real data is not particularly unlikely in any of the toy models they used for comparison. Their models include Gaussian models with an extreme initial power spectrum with power only on scales $\\sim 100h^{-1}$Mpc. They found regular `galaxy' distributions a few per cent of the time and concluded that the BEKS data do {\\it not} rule out all Gaussian models. However, these theoretical studies did not give any clear answer to the question posed above: is the BEKS regularity compatible with the standard CDM paradigm? We attempt to answer this below using versions of all the statistical tools developed in earlier papers. There have been several interesting observational developments after BEKS. Willmer et al. (1994) found that, if the original BEKS deep survey at the North Galactic Pole had been carried out 1 degree or more to the west, many of the peaks would have been missed. On the other hand, Koo et al. (1993) added new data from a wider survey to the original BEKS data and found the highest peak in the power spectrum to be further enhanced. They also analysed another set of deep pencil-beam surveys and found a peak of weaker significance on the {\\it same} scale, 128 $h^{-1}$Mpc. This raises another question: is 128 $h^{-1}$Mpc a preferred length scale for the galaxy distribution? Further support for such a preferred scale has been presented by Tully et al. (1992), Ettori et al. (1997) and Einasto et al. (1997). Thus one can wonder whether a single scale could be indicated with such apparent consistency within the CDM paradigm. With the important exception of the work of PG there has been surprisingly little comparison of the BEKS data with direct simulations of standard CDM cosmogonies. Even before the BEKS discovery, White et al. (1987) had shown that pencil-beams drilled through periodic replications of their CDM simulations frequently showed a kind of `picket fence' regularity in their redshift distribution. Frenk (1991) confirmed this result and concluded that regular patterns similar to that seen in the BEKS data are easy to find in their simulations. However, it is clearly dangerous to make use of periodic replications of a simulation when assessing the significance of apparent periodicities in the redshift distribution. It is preferable to simulate a volume large enough to encompass the whole survey. Furthermore, since many independent artificial surveys are needed to establish that the real data are highly unlikely in the cosmogony simulated, the simulated volume must be fully three-dimensional (unlike that of PG) to allow the creation of many quasi-independent lines-of-sight. A final consideration is that the BEKS data reach to redshifts beyond 0.3, so that evolution of clustering along the survey may not be negligible. In this paper we investigate the distribution of `galaxies' along the past light-cones of hypothetical observers. Particle positions and velocities on these light-cones were generated as output from the Hubble Volume Simulations (Evrard et al. 2000). These very large CDM $N$-body simulations were recently performed by the Virgo consortium and each used $10^9$ particles to follow the evolution of the matter distribution within cubic regions of an $\\Omega=1$ $\\tau$CDM ($\\Omega=0.3$ $\\Lambda$CDM) universe of side 2000 $h^{-1}$Mpc (3000 $h^{-1}$Mpc). Such large volumes allow many independent light-cones to be generated out to $z \\sim 1$. The light-cone output automatically accounts for clustering evolution with redshift. The principal uncertainty lies in how to create a `galaxy' distribution from the simulated mass distribution. We employ Lagrangian bias schemes similar to those of White et al. (1987) and Cole et al. (1998). Individual particles are tagged as galaxies with a probability which depends only on the smoothed {\\it initial} overdensity field in their neighbourhood. The parameters of these schemes are adjusted so that the present-day correlations of the simulated galaxies match observation. Many quasi-independent mock pencil-beam surveys can then be created adopting the geometry and the galaxy selection probability with distance of the BEKS surveys. Our discussions focus primarily on the significance of the BEKS data in comparison with our CDM samples. We begin by following the methods used originally by BEKS, namely, redshift counts, pair separation distributions, and the one-dimensional power spectrum. Redshift counts are translated into a distribution in physical distance assuming the same cosmological parameters as BEKS. For the one-dimensional power spectra, the height of the highest peak is the most important statistic. Szalay et al. (1991) show that the statistical significance of the highest peak of the BEKS data is at $10^{-4}$ level, based on the formal probability for the peak height. This calculation was disputed by Kaiser\\& Peacock (1991) because of the difficulty in estimating the appropriate noise level. We calculate relative peak-to-noise ratios of the highest peaks in the power spectra in identical ways for real and simulated data and so can compare the two without needing to resolve this issue. We also apply two additional statistical tests for regularity, the $\\Delta$ test of PG and a `supercluster' statistic designed by Dekel et al. (1992). Our paper is organised as follows. In section 2 we present details of the $N$-body simulations from which our pencil-beam samples are drawn. In section 3 we explain our bias scheme. In section 4 we describe our mock pencil-beam surveys which mimic as closely as possible the actual observations of BEKS. Our main results for power spectra are given in section 5. Results are given in section 6 for the $\\Delta$ test, and in section 7 for the `supercluster' statistic. We present our conclusions in section 8. ", "conclusions": "By creating a number of mock pencil-beam surveys we have compared the apparent periodicity in two CDM model universes with that observed in the data of Broadhurst et al. (1990). The power spectrum analysis alone shows that the BEKS data are significantly more periodic than the models at about the half per cent level, while the PG $\\Delta$-test shows less significance, about 10 per cent for {\\bf t1} and 5 per cent for {\\bf L1} and {\\bf L2}. The supercluster statistic gives a two per cent probability of finding a structure as regular as the BEKS data. Restricting to a length scale $\\sim$100-150$h^{-1}$Mpc, however, the number of samples which show the kind of periodicity seen in the BEKS data is extremely small for each of these statistics. Overall no sample is more regular than the BEKS data for {\\it all} three statistics for a single period. The two popular CDM models we have studied here are apparently unsuccessful in reproducing the observed periodicity. From this result together with the fact that the statistical results appeared to be insensitive to the choice of the bias model, we conclude that CDM models conflict with the BEKS observation. Either the models need additional physics, or the data are a fluke or are somehow biased. Various possible physical explanations have been proposed, such as coherent peculiar velocities (Hill, Steinhardt and Turner 1991) oscillations in the Hubble parameter (Morikawa 1991) or baryonic features in the power spectrum (Eisenstein \\etal 1998) but all of them seem to require either additional mechanisms with fine tunings beyond the standard theory or cosmological parameters significantly different from currently favoured values. Intriguingly, Dekel et al. (1992) demonstrated that built-in power on a large ($\\sim 100 h^{-1}$Mpc) length scale in the initial density fluctuation could indeed reproduce periodic features on a given scale, at least by some of the tests we have considered. If such excess power on large scales (hence still in the linear regime) exists, it will be detectable in the power spectra of the future 2dF and Sloan surveys. Having found at least a few examples that are nearly as periodic as the BEKS data, we cannot rule out the possibility that the BEKS data (or the Galactic Pole direction) are a fluke. On the other hand, one should be aware of the complexity of the original observations --an incomplete compilation of a narrow and deep, and of a wide and shallow survey at each of the Galactic Poles. It is not clear whether such a combination constitutes a fair sample. No evidence for periodic structure on $\\sim$130$h^{-1}$Mpc has been found so far in two other deep redshift surveys, the ESO-Sculptor Survey (Bellanger and de Lapparent 1995) and the Caltech Faint Galaxy Redshift Survey (Cohen 1999). Follow-up observations to BEKS by Koo \\etal (1993) did {\\it not} show a strong regularity in two other directions, although around the Galactic Pole the regularity was found to be further strengthened. Our results give the {\\it a priori} probability for such apparent periodicity in CDM models. Several more deep surveys might suffice to judge whether the discrepancy with BEKS reflects a major inconsistency. The planned VIRMOS Deep Survey (Le F\\`{e}vre \\etal 1998, see also Guzzo 1999) will survey the range $0.3\\le z\\le 1$ and will provide, together with the large volume 2dF and Sloan surveys, much larger and more complete samples in the near future." }, "0011/hep-ph0011097_arXiv.txt": { "abstract": "One of the possible consequences of the existence of extra degrees of freedom beyond the electroweak scale is the increase of neutrino-nucleon cross sections ($\\sigma_{\\nu N}$) beyond Standard Model predictions. At ultra-high energies this may allow the existence of neutrino-initiated extensive air showers. In this paper, we examine the most relevant observables of such showers. Our analysis indicates that the future Pierre Auger Observatory could be potentially powerful in probing models with large compact dimensions. ", "introduction": " ", "conclusions": "" }, "0011/astro-ph0011107_arXiv.txt": { "abstract": "Relativistic kinetic theory, based on the Grad method of moments as developed by Israel and Stewart, is used to model viscous and thermal dissipation in neutron star matter and determine an upper limit on the maximum mass of neutron stars. In the context of kinetic theory, the equation of state must satisfy a set of constraints in order for the equilibrium states of the fluid to be thermodynamically stable and for perturbations from equilibrium to propagate causally via hyperbolic equations. Application of these constraints to neutron star matter restricts the stiffness of the most incompressible equation of state compatible with causality to be softer than the maximally incompressible equation of state that results from requiring the adiabatic sound speed to not exceed the speed of light. Using three equations of state based on experimental nucleon-nucleon scattering data and properties of light nuclei up to twice normal nuclear energy density, and the kinetic theory maximally incompressible equation of state at higher density, an upper limit on the maximum mass of neutron stars averaging 2.64 solar masses is derived. ", "introduction": "The problem of the maximum possible mass of neutron stars has a long history. Baade and Zwicky in 1934 proposed a star composed of densely packed neutrons as the final state resulting from the supernova process \\cite{Baade34}. Oppenheimer and Volkoff in 1939 demonstrated a star composed of noninteracting neutrons is supported against gravitational collapse to a black hole by the Fermi degeneracy pressure only for stellar masses up to 0.72 solar masses ($M_\\odot$) \\cite{Oppenheimer39}. The discovery of the first pulsar in 1967 and the realization pulsars are highly magnetized rotating neutron stars triggered substantial theoretical work towards understanding the structure of neutron stars. Pulsar masses can be measured from observations of Doppler shifts of periodic emissions from neutron stars in binary systems. Some 20 radio pulsar masses have been accurately determined so far, with most masses clustering around 1.4 $M_\\odot$ \\cite{Thorsett99}. This is well beyond the value of 0.72 $M_\\odot$ computed for noninteracting neutrons, so the short-range repulsion of the strong nuclear force must make a substantial contribution to the pressure supporting neutron stars against gravitational collapse. There is also evidence for neutron stars with substantially larger masses than the radio pulsars due to accretion of matter from a binary companion. The mass of the x-ray pulsar in the Vela X-1 binary has been estimated to be 1.9 $M_\\odot$ \\cite{vanKerkwijk00}, and the x-ray burster Cygnus X-2 has been determined to be 1.8 $M_\\odot$ \\cite{Orosz00}. Quasiperiodic oscillations in the x-ray emissions from neutron stars accreting matter from low mass companions have been argued to imply a mass of up to 2.3 $M_\\odot$ for the neutron stars in these systems \\cite{Zhang97,Miller98}. An accurate theoretical determination of the maximum possible mass of neutron stars is of practical interest for identifying as a black hole any compact object with a larger mass. The maximum mass value remains uncertain because the equation of state of neutron star matter is not well understood at the high density values found in neutron star interiors. Modern methods develop the equation of state by fitting experimental nucleon-nucleon scattering data and properties of light nuclei to two and three-body interaction potentials \\cite{Wiringa88,Akmal98}. Confidence in the results of these methods is high near normal nuclear density (energy density = 152 MeV/fm$^3$ = 2.7 X 10$^{14}$ g/cm$^3$, baryon density = 0.16 fm$^{-3}$) because, at low density, experimental verification is possible with existing facilities. The validity of extrapolating the experimentally verified low density equation of state to high density, or as an alternative theoretically modeling the physical processes thought to occur at high density, is uncertain because of the lack of high density laboratory data. The sources of the uncertainties in the high density equation of state are incomplete knowledge of the three-nucleon interactions, the contributions of mesonic and other baryonic besides nucleonic degrees of freedom, and a possible quark deconfinement transition \\cite{Akmal98}. In light of these uncertainties in the exact form of the high density equation of state, several authors have derived an upper limit on the maximum neutron star mass using only general restrictions on the equation of state. Oppenheimer and Volkoff suggested \\cite{Oppenheimer39}, and Rhoades and Ruffini rigorously proved \\cite{Rhoades74}, the maximum mass of a stable neutron star results when the stiffest equation of state compatible with thermodynamic stability and causality is used. Rhoades and Ruffini derived an upper limit on the maximum neutron star mass of 3.2 $M_\\odot$ by using the experimentally verified Harrison-Wheeler equation of state \\cite{Harrison65} up to 1.7 times normal nuclear energy density, and the stiffest equation of state compatible with causality at high density. They defined a casual equation of state to be one having an adiabatic sound speed less than the speed of light, so they took the most incompressible equation of state to have the sound speed equal to light speed. Kalogera and Baym \\cite{Kalogera96} updated the result of Rhoades and Ruffini, using an equation of state developed by Wiringa, Fiks, and Fabrocini \\cite{Wiringa88} up to twice normal nuclear energy density. They derived an upper mass limit of 2.9 $M_\\odot$ using the same maximally incompressible equation of state as Rhoades and Ruffini above twice normal nuclear energy density. The purpose of this paper is to apply insights from relativistic kinetic theory to the problem of the maximum possible mass of neutron stars. Kinetic theory used to model viscosity and heat conduction shows requiring the sound speed to be less than the speed of light is a necessary but not sufficient condition for causality. The equation of state must satisfy a set of constraining conditions for the fluid to have stable equilibrium states with perturbations from equilbrium that propagate causally via a system of hyperbolic equations \\cite{Hiscock83,Olson90}. If any one of these constraints is violated, there will be at least one fluid mode with a superluminal propagation speed. In particular, it is possible for the fluid to be acausal in a transverse mode or a different longitudinal mode even though the sound speed is below light speed. These constraints generally fix the stiffest possible equation of state to have an adiabatic sound speed significantly less than the speed of light. The main result of the present paper is application of these constraints to neutron star matter at high density softens the maximally incompressible equation of state and reduces the maximum neutron star mass significantly below the value found using a high density equation of state having the adiabatic sound speed equal to the speed of light. In the next section the thermodynamic constraint conditions from relativistic kinetic theory are reviewed. Then the maximally incompressible equation of state of neutron star matter is developed from these constraint conditions. Finally, neutron star models are constructed using this stiffest possible equation of state at high density, and the resulting maximum neutron star mass is determined. Gravitational units having $G=c=1$ are used. ", "conclusions": "The principal finding of this paper is stability and causality constraints resulting from applying kinetic theory to thermal and viscous dissipative processes in neutron star matter yields a softer maximally incompressible equation of state, and a lower maximum mass of neutron stars, than for the sound speed equal to light speed maximally incompressible equation of state. The reduction in the maximum mass averages from 2.93 $M_\\odot$ down to 2.64 $M_\\odot$ for the three low density equations of state considered if the low density equation of state is regarded as experimentally verified up to twice normal nuclear energy density. The softening of the maximally incompressible equation of state is a generic prediction of kinetic theory, but the specific value of the resulting maximum neutron star mass derived here rests upon the accuracy of approximating the relaxation times and viscous-thermal couplings with their noninteracting degenerate Fermi gas forms, and the overall validity of dissipative relativistic kinetic theory for strongly interacting nuclear matter at high density. Future data from the Relativistic Heavy Ion Collider holds promise for experimental testing of the applicability of the kinetic theory approach to modeling high density nuclear matter. Only static neutron star mass models have been constructed in this paper. Rotation can support larger mass stars against collapse, but only for stars rotating rapidly enough to be nearly shedding mass from the equator is the maximum mass value significantly changed from its nonrotating value. For these rapidly rotating stars the increase in the maximum mass is at most approximately 20\\% over the static value when undergoing uniform rotation \\cite{Cook92,Cook94}. Differential rotation is strongly damped \\cite{Hegyi77}. The remnant of binary neutron star coalescence may undergo a short period of dynamically stable differential rotation, and thus briefly support a much larger mass neutron star remnant against prompt black hole collapse than is possible for a static or uniformly rotating star \\cite{Baumgarter00}." }, "0011/astro-ph0011331_arXiv.txt": { "abstract": "We have used the BIMA array to image the Galactic Center with a 19-pointing mosaic in HCN(1-0), HCO$^+$(1-0), and H 42$\\alpha$ emission with 5 km s$^{-1}$ velocity resolution and $13'' \\times 4''$ angular resolution. The $5'$ field includes the circumnuclear ring (CND) and parts of the 20 and 50 km s$^{-1}$ clouds. HCN(1-0) and HCO$^+$ trace the CND and nearby giant molecular clouds while the H 42$\\alpha$ emission traces the ionized gas in Sgr A West. We find that the CND has a definite outer edge in HCN and HCO$^+$ emission at $\\sim45''$ radius and appears to be composed of two or three distinct streams of molecular gas rotating around the nucleus. Outside the CND, HCN and HCO$^+$ trace dense clumps of high-velocity gas in addition to optically thick emission from the 20 and 50 km s$^{-1}$ clouds. A molecular ridge of compressed gas and dust, traced in NH$_3$ emission and self-absorbed HCN and HCO$^+$, wraps around the eastern edge of Sgr A East. Just inside this ridge are several arcs of gas which have been accelerated by the impact of Sgr A East with the 50 km s$^{-1}$ cloud. HCN and HCO$^+$ emission trace the extension of the northern arm of Sgr A West which appears to be an independent stream of neutral and ionized gas and dust originating outside the CND. Broad line widths and OH maser emission mark the intersection of the northern arm and the CND. Comparison to previous NH$_3$ and 1.2mm dust observations shows that HCN and HCO$^+$ preferentially trace the CND and are weaker tracers of the GMCs than NH$_3$ and dust. We discuss possible scenarios for the emission mechanisms and environment at the Galactic center which could explain the differences in these images. ", "introduction": "A 2-5 pc circumnuclear ring (CND) of neutral gas and dust has been imaged at the Galactic Center in the far-infrared and radio \\citep{gen85,gus87,mar93}. It has been proposed that the CND is part of an accreting disk of material surrounding the massive central source Sgr A$^*$ \\citep{gus87}, an apparent supermassive black hole \\citep{ghe98,gen97}. Ionized gas streamers appear to be infalling towards the central mass, originating at the inner edge of the CND and converging at the location of the bar near Sgr A$^*$ \\citep{loe83}. It is unclear whether the CND is a transitory or stable structure. If it is stable, the question remains as to how the CND itself is maintained. At a projected distance of 10 pc from the Galactic Center lie two giant molecular clouds (GMC), the 20 km s$^{-1}$ cloud (M-0.13-0.08) and the 50 km s$^{-1}$ cloud (M-0.02-0.07) \\citep{gus81} to the south and east of the CND respectively. \\citet{gus81b} argued that these clouds must lie more than 100 pc from the Galactic Center in order not to be tidally disrupted. Their interpretation of formaldehyde absorption data places the 50 km s$^{-1}$ cloud behind the nucleus and the 20 km s$^{-1}$ cloud in front. Various studies have investigated the interactions between these molecular clouds, the non-thermal radio source Sgr A East, the CND and the ionized gas streamers in the nucleus \\citep{hoe85, gen90, hoe91, ser92, den93, mar95, coi00}. There are suggestions of molecular gas steamers from both of the GMCs located 10 pc away in projection that may be feeding the CND. A southern streamer has been mapped in NH$_3$(1,1), (2,2) and (3,3) and sub-millimeter continuum \\citep{hoe91, den93, coi99} and emission in the same region has been seen in HCN(3--2) \\citep{mar95}. An eastern streamer, lying at the edge of the Sgr A East SNR and protruding in towards the nuclear region, has been mapped in HCN(1--0) \\citep{hoe93} and HCN(4--3) and HCN(3--2) \\citep{mar95}. This streamer has not been seen in NH$_3$ or sub-millimeter continuum. More recently, \\citet{coi00} presented images of two streams of NH$_3$(1,1) and (2,2) emission parallel to the Galactic plane, which they interpret as evidence for infall from the 20 km s$^{-1}$ cloud onto the CND, and as evidence for interactions between Sgr A East and the 50 km s$^{-1}$ cloud, and a SNR to the south of Sgr A East (G359.92--0.09). They argue that both clouds are within 10 pc of the nucleus and that only the 20 km s$^{-1}$ cloud is interacting with the CND. Many of these observations have a limited field-of-view and velocity coverage, mapping only the central $2-4'$ (5--10pc) of the Galaxy and not extending into the large molecular clouds, thereby unable to thoroughly address the question of possible interactions with the circumnuclear region. This paper describes observations with the BIMA array\\footnote{The BIMA array is operated by the Berkeley-Illinois-Maryland Association under funding from the National Science Foundation.} of the central $5'$ in HCO$^+$, HCN(1-0), and H 42$\\alpha$ emission with 5 km s$^{-1}$ resolution, and in 87 GHz continuum emission. The $5'$ field of view corresponds to a field roughly 12.5 pc at the distance of the Galactic Center (assuming $R_\\odot$ = 8.5 kpc). ", "conclusions": "i) We present images of HCN(1-0) and HCO$^+$ emission and absorption which trace dense molecular gas in the central 12 parsecs of the Galaxy. The HCN(1-0) and HCO$^+$ emission primarily traces filamentary structures associated with the CND and nearby GMCs; low-level, extended emission from partially-resolved larger structures is also seen. ii) The CND appears as a bright, well-defined ring of emission $20'' - 60''$ from the center, at an inclination of 60--70$^{\\circ}$. Although the outer edge of the ring is not as sharply defined as the inner edge, the CND does not extend beyond $100''$ in HCN(1-0) and HCO$^+$ emission. The CND does not appear to be an equilibrium structure; rather, it consists of two or three separate streamers in rotation around the nucleus. The CND may exist long-term in a fluctuating sense, as gas streamers feed the inner parsec. iii) Outside the CND, HCN(1-0) and HCO$^+$ trace several filamentary structures which do not follow the rotation pattern of the $45''$ ring. High-velocity emission is seen $90''$ east of Sgr A*, just inside the dense molecular ridge of shocked material seen at the intersection of Sgr A East and the 50 km sec$^{-1}$ cloud. A stream of emission is also seen extending from the ionized northern arm, through the CND to high-velocity red- and blue-shifted material to the north. This material may be falling into the center on an orbit which intersects the CND streamers. iv) Comparison of the HCN(1-0) and HCO$^+$ with NH$_3$ and dust emission shows that NH$_3$ emission is well correlated with self-absorbed HCN(1-0) and HCO$^+$ emission, and dust ridges in the nearby GMCs. v) HCN(1-0) and HCO$^+$ emission appear to be enhanced in the CND relative to NH$_3$. Further observations of other spectral lines are required to distinguish between opacity, excitation and chemical explanations for these differences. These observations show that the accretion of material into the Galactic Center region may be a chaotic and transient phenomenon. From the spatial distribution of the nearby gas with respect to Sgr A East and other nearby supernova remnants, and from the overall kinematics and especially the isolated pockets of high-velocity gas, we have the impression that the nearby molecular clouds are being disrupted by the expanding shells associated with Sgr A East. Infall of material appears to approach the CND from a number of different directions on different orbits. Sgr A East itself may be driven by multiple outflow events associated with the central black hole or multiple supernovae. We favor a model where the accretion process is highly non-steady and non-uniform at least on the timescale of the rotation period of the CND, 10$^{5}$ yrs. In this scenario, the CND does not appear uniform because material continues to fall into the CND on short timescales. Whether the further infall of material from the CND toward the central black hole is also transient and non-steady in nature is unclear from these studies. However, there is some evidence that material may approach the central region directly from outside of the CND via individual streamers. {\\it Acknowledgments} This work was supported in part by NSF Grant AST-21795 to the University of California. ALC is supported by an NSF Graduate Research Fellowship. We thank Robert Zylka for permission to use the 1.2 mm image. \\newpage" }, "0011/astro-ph0011041_arXiv.txt": { "abstract": "N-body + SPH simulations are used to study the evolution of dwarf irregular galaxies (dIrrs) entering the dark matter halo of the Milky Way or M31 on plunging orbits. We propose a new dynamical mechanism driving the evolution of gas rich, rotationally supported dIrrs, mostly found at the outskirts of the Local Group (LG), into gas free, pressure supported dwarf spheroidals (dSphs) or dwarf ellipticals (dEs), observed to cluster around the two giant spirals. The initial model galaxies are exponential disks embedded in massive dark matter halos and reproduce nearby dIrrs. Repeated tidal shocks at the pericenter of their orbit partially strip their halo and disk and trigger dynamical instabilities that dramatically reshape their stellar component. After only 2-3 orbits low surface brightness (LSB) dIrrs are transformed into dSphs, while high surface brightness (HSB) dIrrs evolve into dEs. This evolutionary mechanism naturally leads to the morphology-density relation observed for LG dwarfs. Dwarfs surrounded by very dense dark matter halos, like the archetypical dIrr GR8, are turned into Draco or Ursa Minor, the faintest and most dark matter dominated among LG dSphs. If disks include a gaseous component, this is both tidally stripped and consumed in periodic bursts of star formation. The resulting star formation histories are in good qualitative agreement with those derived using HST color-magnitude diagrams for local dSphs. ", "introduction": "Dwarf galaxies in the Local Group (LG) clearly obey a morphology-density relation. Close to the Milky Way and M31 we find early-type dwarf galaxies, namely faint ($M_B > -14$) low surface brightness dwarf spheroidals (dSphs) and more luminous ($M_B > -17$), higher surface brightness dwarf ellipticals (dEs). All these galaxies are nearly devoid of gas, contain dark matter and mainly old stars and are supported by velocity dispersion (Ferguson \\& Binggeli 1994, hereafter FB94; Mateo 1998, hereafter Ma98; Grebel 1999, hereafter Gr99; Van den Bergh 1999). Among them Draco and Ursa Minor have the highest dark matter densities ever measured (Lake 1990). On the outskirts of the LG we find similarly faint ($M_B>$ -18) and dark matter dominated dwarf irregular galaxies (dIrrs), that are gas rich, star-forming systems with disk-like kinematics (Ma98, Van den Bergh 1999, Gr99). Previous attempts to explain the origin of dSphs in the LG have relied on gas dynamical processes to remove the gas in dIrrs. Gas stripping may result either because of the pressure exerted by an external hot gaseous medium in the halo of the Milky Way (``ram pressure'') (Einasto et al. 1974) or because of internal strong supernovae winds (Dekel \\& Silk 1986). However, ram pressure would require an external gas density that is several orders of magnitude higher than recently inferred for the Milky Way (Murali 2000) and supernovae winds cannot explain the existing morphology-density relation. Moreover, such dissipative mechanisms would remove the gas but would not directly alter the structure and kinematics of the pre-existing stellar component. However, the light follows an exponential profile in both dSphs and dIrrs (Faber \\& Lin 1983; Irwin \\& Hatzdimitriou 1995; Ma98) and a positive correlation between surface brightness and luminosity is shown by both types of dwarfs (FB94), suggesting an evolutionary link between them. Is there a mechanism that can transform dwarf galaxies between morphological classes or must we accept the idea that dSphs are fundamentally different from dIrrs ? Within rich galaxy clusters, fast fly-by encounters with the largest galaxies can transform a disk system into a spheroidal or S0 galaxy in just 3-4 Gyr (Moore et al. 1996, 1998). If the halos of bright galaxies were scaled down versions of galaxy clusters then this ``galaxy harassment'' would be equally important within them. However, whereas rich clusters contain over thirty large ($L_*$) perturbing galaxies, the Milky Way and M31 have only a couple of satellites sufficiently massive to harass the other dwarf galaxies (Moore et al. 1999; Klypin et al. 1999) As a result, the rate for effective satellite-satellite fly-by encounters is less than one in every 10 Gyr (the LMC and the SMC being a notable exception). Thus, we are left only with the repeated action of tidal forces from the primary galaxy as an evolutionary driver. These operate on the orbital timescale, which is of order of 3-4 Gyr in both clusters and galactic halos. However, given the relatively low age of large, virialized clusters, galaxies have typically approached the cluster center only once by the present time, while dSphs satellites have had sufficient time to complete several close tidal encounters with the Milky Way, as stellar ages imply that the latter was already in place 10 Gyr ago (Van den Bergh 1996). In this paper we use very high resolution N-Body + SPH simulations performed with the parallel binary treecode GASOLINE (Dikaiakos \\& Stadel 1996; Wadsley et al. 2000) to follow the evolution of small galaxies resembling dIrrs as they move on bound orbits in the tidal field of the massive dark matter halo of the Milky Way. ", "conclusions": "Figure 5 summarizes the main observable properties of the simulated satellites projecting them on the Fundamental Plane (FB94). The remnants of LSB satellites resemble dSphs like Fornax or Sagittarius ($-14 < M_B < -11$), while HSBs transform in the bright dEs ($M_{B} > -17$), having a final central surface brightness higher than that of observed dIrrs with the same luminosities and therefore matching another observational constraint (FB94;Ma98). The total (including dark matter) final mass-to-light ratios are in the range $6-20$. Remarkably, our model can reproduce the properties of even the most extreme dSphs, Draco and Ursa Minor. In fact, as the dark matter halo of GR8 is barely affected by tides (Figure 1), the remnant ($M_B = -7.5$) has a final mass-to-light ratio still $\\sim 50$ and the central dark matter density is still around $0.3 M_{\\odot}$ pc$^{-3}$, matching the structural parameters inferred for Draco and Ursa Minor (Ma98). \\medskip \\epsfxsize=8truecm \\epsfbox{fundpnew.ps} \\figcaption[fundpnew.ps]{\\label{fig:asymptotic} \\small{Fundamental Plane (FP) for all the remnants as as in Ferguson \\& Binggeli (1994) (bottom). In the FP plot, $\\mu$ and $\\sigma$ are, respectively, the average surface brightness and velocity dispersion measured inside $R_e$ and $M_B$ is measured at the Holmberg radius. The dashed line is a fit to the distribution of local dSphs, while the solid line refers to elliptical galaxies within the Virgo cluster (Dressler et al. 1987). Open and filled symbols represent, respectively, the initial and final state of HSBs (squares) and LSBs (circles). The initial and final state of the GR8 model are indicated by the open triangle and ``star''.}} ``Tidal stirring'' naturally leads to the spatial segregation of dIrrs versus dSph as its effectiveness depends strongly on the distance from the primary. How important is our assumption of a massive and extended dark matter halo surrounding the Milky Way ? When we adopt a ``minimal'' dark halo truncated at 50 kpc (with mass $5 \\times 10^{11} M_{\\odot}$; Little \\& Tremaine 1987), tides are too weak and the final remnants are still rotationally flattened ($v/\\sigma > 1$). Instead, within a halo as massive and extended as implied by theories of galaxy formation (Peebles 1989) , our dIrrs models transform into dSphs even on orbits with apocenters larger than 200 kpc, explaining the origin of even the farthest dSphs as Leo I and Leo II. Though rather speculative at this stage, it is tempting to relate HSB satellites observed during the strong bursting phase to the population of blue compact dwarfs identified by Guzm\\'an et al. (1997) at intermediate redshift. Redshift surveys will establish if bursting dwarfs have nearby massive companions. Extended tidal streams of stars originate from our simulated dwarfs (Fig. 2) with a maximum surface brightness of just 30 mag arcsec$^{-2}$ (B band). Spectroscopic evidence for stellar streams from the dSph Carina has been recently claimed (Majewski et al. 2000). Future astrometric missions, like SIM and GAIA (Gilmore et al. 1998; Helmi et % should reveal such faint features and will also carry out high-quality measurements of proper motions for many satellites of the Milky Way, thus providing a test for the orbital configurations used in this model. Our model successfully explains the origin of dSphs once all observational constraints are taken into account: they evolved from dIrrs that entered the halo of the Milky Way or M31 several Gyr ago moving on plunging orbits and suffered stirring by the tidal field of the large spirals. \\vskip 16 pt The authors thank G.Bothun for stimulating discussions. Simulations were carried out at the CINECA (Bologna) and ARSC (Fairbanks) supercomputing centers." }, "0011/astro-ph0011277_arXiv.txt": { "abstract": "\\noindent From theoretical predictions, the introduction of 8--m class telescopes permits one to extend Surface Brightness Fluctuations measurements from the ground to $\\approx$7000 km/s, with a precision that is comparable to current space--based measurements. We have measured I--band SBF in IC 4296 in Abell 3565 $cz \\sim 3630$ km/s with the ESO Very Large Telescope. Adopting the Tonry et al. 2000 calibration for I--band SBF we determined a distance modulus of $(\\overline I_{o,k} - \\overline M_{I}) = 33.44 \\pm 0.17$ corresponding to a galaxy distance of 49 $\\pm$ 4 Mpc. This result is consistent with the HST observation from Lauer et al. 1998: $(\\overline I_{F814} - \\overline M_{F814}) = 33.47 \\pm 0.13$. ", "introduction": "\\noindent I--band Surface Brightness Fluctuations (SBF) have been successfully used to measure early-type galaxy distances up to 4000 km/s from the ground and up to 7000 km/s with the Hubble Space Telescope \\cite{sod95,sod96,aj97,ton97,tho97,lau98,pah99,bla99a,ton00}. This method was introduced by Tonry \\& Schneider 1988 (a recent review has been given by Blakeslee et al. 1999a) and is based on a simple concept. The Poissonian distribution of unresolved stars in a galaxy produces fluctuations in each pixel of the galaxy image. The variance of these fluctuations is inversely proportional to the square of the galaxy distance. The SBF amplitude is defined like this variance normalized to the mean flux of the galaxy in each pixel \\cite{ts88}. The stars that contribute the most to the SBF are then the brightest stars in the intrinsic luminosity function, typically red giants in old stellar populations. The absolute magnitude of the fluctuation is not a constant, but depends on the age and metallicity of the stellar population. As such it is a strong function of not only the photometric band in which the observations are carried out, but also of the color of the galaxy under study. Tonry et al. (1997) and Tonry et al. (2000) have quantified these dependencies using an extensive sample of I and V band observations, which are used to empirically calibrate the dependence of the I-band fluctuation magnitude on (V-I) color. The absolute calibration is then obtained using a set of 5 galaxies with independent Cepheid distance. The I--band SBF absolute magnitude calibration can be written as: \\begin{equation} \\overline M_I=-1.74+4.5\\ [(V-I)_0-1.15]. \\label{eq:Ical} \\end{equation} \\noindent Ferrarese et al. (2000a) have obtained their own calibration based on accurate Cepheid distances to six spiral galaxies with SBF measurements within 1200 km/s. The Cepheid distances were part of a larger dataset of 18 distances to galaxies within the Fornax and Virgo clusters observed by the HST Key Project on the Extragalactic Distance Scale. The two calibrations are consistent within the errors. \\small{ \\begin{table*}[!htf] \\begin{flushleft} \\begin{tabular} {|c|c|c|c|c|c|c|c|c|} \\hline Name&RA &DEC &l&b&Type&$m_V$&$V-I$&$V_{LG} $ \\\\ \\hline &(2000)&(2000)°°&&mag&mag&km/s \\\\ \\hline IC 4296&13 36 39.46 & -33 57 59.8 &313.54 &+27.97&E&10.57&1.24&3630 \\\\ \\hline \\end{tabular} \\end{flushleft} \\caption{General properties of IC 4296, from the web archive Simbad (http://simbad.u-strasbg.fr), and Lauer et al. 1998} \\label{tab-vlt} \\end{table*} } \\normalsize \\noindent The potential of these measurements extends to the measurements of cosmological parameters and the study of the dynamics of the universe. SBF distances have been used in measuring the Hubble constant in the HST Key project on the Extragalactic Distance Scale (Ferrarese et al. 2000a, and references therein). Tonry and his collaborators \\cite{ton00} have derived peculiar velocities from their ground based sample up to 4000~km/s and calculated bulk flows in this region. By comparing these peculiar velocities with infrared and optical surveys, Blakeslee et al. 1999a have been able to constrain the value of the Hubble constant and the parameter density (luminous plus dark matter) $\\beta=\\Omega^{0.6}/b$, where $b$ is the linear bias \\cite{bla99b}. \\noindent However, the current I--band sample is limited to 4000~km/s. The introduction of large telescopes, such as the Very Large Telescope (VLT), have opened new opportunities to extend the current sample from the ground. With the optical instrument FORS1 and FORS2 on VLT, from a theoretical point of view, it will be possible to extend I--band SBF with high signal--to--noise and good understanding of external source contribution up to $\\approx$ 7000~km/s \\cite{mei00}. This implies that ground--based measurements can be used to probe the same volume as HST measurements. \\noindent To illustrate the potential of I--band SBF measurements with the VLT, we have observed IC 4296 in Abell 3565. This galaxy is part of the brightest ellipticals in the Abell clusters observed in the Lauer \\& Postman 1992 sample. An I--band SBF distance measurement of this galaxy, based on HST observations, has already been reported by Lauer et al. (1998). We compare our SBF measurements with Lauer et al. (1998) and obtain similar results and measurement precision, being our errors and external source contribution estimation comparable with HST capabilities. ", "conclusions": "\\noindent I--band SBF have been observed in IC 4296 with the VLT. A distance modulus of $(\\overline I_{o,k} - \\overline M_I) = 33.44 \\pm 0.17$ was measured, from which a galaxy distance $49 \\pm 4$ Mpc is derived. \\noindent This result confirm the potential of 8--m class telescopes in these kind of measurements and suggest that future SBF observations from the ground can reach the same distance that until now were only reachable by HST." }, "0011/astro-ph0011457_arXiv.txt": { "abstract": "s{ The evolution of the X-ray luminosity function of clusters of galaxies has been measured to z=0.9 using over 150 X-ray selected clusters discovered in the WARPS survey. We find no evidence for evolution of the luminosity function at any luminosity or redshift. The observations constrain the evolution of the space density of moderate luminosity clusters to be very small, and much less than predicted by most models of the growth of structure with $\\Omega_m$=1. All the current X-ray surveys agree on this result. Several notable luminous clusters at z$>$0.8 have been found, including one cluster which is more luminous (and is probably more massive) than the well known MS1054 cluster. } ", "introduction": "The evolution of the space density of clusters of galaxies is a measurement sensitive to the physical and cosmological parameters of models of structure formation. We describe a deep X-ray survey of clusters of galaxies (the Wide Angle ROSAT Pointed Survey or WARPS), and use it to measure the evolution of the X-ray luminosity function (XLF) of clusters of galaxies (Scharf et al 1997, Jones et al 1998, Ebeling et al 2000a, Fairley et al 2000). ", "conclusions": "\\subsection{Comparison with results of X-ray cluster surveys} Several recent deep X-ray clusters surveys are described in these proceedings and elsewhere. In the regime where these surveys have good statistical accuracy (ie. moderate X-ray luminosities $\\sim$10$^{44}$ erg s$^{-1}$) there is excellent agreement that {\\it no evolution of the XLF is observed to z$\\approx$0.8}. Five surveys agree on this point (EMSS, RDCS, SHARC, CfA, WARPS) and the only disagreement (the RIXOS survey of Castander et al 1995) can be understood in terms of the RIXOS source detection algorithm. At the higher X-ray luminosities of the most massive clusters (L$_X>$5x10$^{44}$ erg s$^{-1}$), there is some disagreement between the results of different surveys as to the degree of evolution found at z$>$0.3 (if any). This may partly be due to the small numbers of high luminosity clusters in any one survey. The range of evolution found is not large: from none to negative evolution of a factor $\\approx$3. The WARPS and RDCS (Borgani et al, these proceedings) are both consistent with no evolution of the XLF up to z=1. WARPS is not sensitive to clusters at z$>$1 because our optical followup does not extend into the NIR. \\subsection{Revisiting the EMSS survey} In an effort to understand the EMSS results of Gioia et al (1990) and Henry et al (1992), who found negative evolution at high luminosities but relatively low redshifts (z$\\approx$0.33), we have remeasured the X-ray luminosities of the 11 EMSS clusters at z$>$0.3 for which deep ROSAT PSPC data exists, extending the work started by Jones et al (1998). If the X-ray luminosities of EMSS z$>$0.3 clusters have been significantly underestimated by a factor of $\\approx$2, then the EMSS XLF will move toward a no-evolution result. We use large (3 Mpc radius, H$_0$=50) apertures, removing point sources and using ASCA temperatures where necessary. We find that the mean ratio $L_{X,PSPC}/L_{X,EMSS}\\approx1.6$ and that the ratio is correlated with the core radius measured from the PSPC images. This suggests that the assumption of a constant core radius of 250 kpc in the EMSS (a reasonable assumption, given the data available at the time) has led to an underestimate of the luminosities of clusters with larger core radii. Henry (2000) has investigated EMSS cluster luminosities and notes that the original EMSS extrapolation of the surface brightness profile to give the total flux (a mean correction factor of $\\approx$2.5) did not take into account the Einstein IPC psf. Including the effect of the psf revises the EMSS luminosities upwards by a factor of 1.37, explaining a major part of the discrepancy with the ROSAT luminosities. However, Henry (2000) also notes that luminosities measured with ASCA are only 17\\% higher than EMSS luminosities. \\subsection{Future work} Future work will concentrate on the handful of clusters in the WARPS survey at z$\\approx$0.8, on measurements of the evolution of the cluster X-ray temperature function using Chandra and XMM, and studies of galaxy evolution in X-ray selected clusters. We are also studying the WARPS sample of `fossil' groups of galaxies, to help understand the formation of luminous elliptical galaxies. New wide area X-ray surveys, designed to detect the rare high luminosity, massive clusters in large numbers at high redshifts (z$>$0.5) are planned (see Lumb \\& Jones 2000 and Ebeling et al 2000c). It is the most massive clusters which have the greatest leverage to constrain cosmological parameters." }, "0011/astro-ph0011396_arXiv.txt": { "abstract": "It is shown that, under sufficiently intense OB-star illumination of a stationary photoexcitation front (PDR), nonlinear H$_{2}$ photoexcitation processes comprising driven resonant two-photon transitions between X-state quantum levels, with VUV continuum light from the star supplying both driving fields, largely determine the photonic pathways of H$_{2}$ molecules in the PDR close to the ionization front. Specifically, for a flux of $% \\sim $4 x 10$^{5}$ Habing fields incident upon a PDR, the total rate at which an H$_{2}$ molecule is nonlinearly photoexcited out of any X-state quantum level is calculated to be roughly 100 times greater than the total rate at which it is linearly photoexcited out of the same level. In strongly excited PDRs, the populations in almost all of the $\\sim$300 bound quantum levels of the X state will be maintained approximately equal via a few myriads of interconnecting two-photon steps. The remarkable importance of two-photon transitions in H$_{2}$ photoexcitation in strongly irradiated PDRs derives from the exceptionally narrow Raman linewidth ($\\Gamma \\sim $10$^{-6}$ $\\sec ^{-1}$) that characterizes all two-photon transitions between bound H$_{2}$ X-state quantum levels. ", "introduction": "In existing theoretical models of stationary photoexcitation fronts (PDRs) \\setlength{\\baselineskip}{0.5\\normalbaselineskip} $[$see Draine \\& Bertoldi (1996) for a comprehensive list of references$]$, a one-dimensional geometry is usually considered, with light from an OB-type illuminating star assumed incident upon the PDR from one direction. In all published PDR models, only linear H$_{2}$ photoexcitation processes have been considered. The consensus among astrophysicists is that these models adequately describe the main photonic events that occur in H$_{2}$% -containing clouds separated by at least 10 pc from OB-type illuminating stars. In the present Letter, we draw attention to the fact that, in very strongly irradiated PDRs (for example, those with cloud-to-star distances $% \\leq $0.1 pc), nonlinear photoexcitation of H$_{2}$ can evidently become the dominant photonic process determining the structure of the PDR. ", "conclusions": "" }, "0011/astro-ph0011169_arXiv.txt": { "abstract": "We report the presence of a 106-day cycle in the radio variability of Sgr A* based on an analysis of data observed with the Very Large Array (VLA) over the past 20 years. The pulsed signal is most clearly seen at 1.3 cm with a ratio of cycle frequency to frequency width $f/\\Delta f= 2.2\\pm0.3$. The periodic signal is also clearly observed at 2 cm. At 3.6 cm the detection of a periodic signal is marginal. No significant periodicity is detected at both 6 and 20 cm. Since the sampling function is irregular we performed a number of tests to insure that the observed periodicity is not the result of noise. Similar results were found for a maximum entropy method and periodogram with CLEAN method. The probability of false detection for several different noise distributions is less than 5\\% based on Monte Carlo tests. The radio properties of the pulsed component at 1.3 cm are spectral index $\\alpha\\sim1.0\\pm 0.1$ (for $S\\propto\\nu^\\alpha$), amplitude $\\Delta S=0.42\\pm 0.04 {\\rm\\ Jy}$ and characteristic time scale $\\Delta t_{FWHM}\\approx 25\\pm5$ days. The lack of VLBI detection of a secondary component suggests that the variability occurs within Sgr A* on a scale of $\\sim5$ AU, suggesting an instability of the accretion disk. ", "introduction": "The compact radio source Sagittarius A* is suggested to be associated with a supermassive black hole at the Galactic Center (Eckart \\& Genzel \\etal 1997, Ghez \\etal 1998, Backer \\& Sramek 1999, Reid \\etal 1999). The flux density variability of Sgr A* has been puzzling since the discovery of this intriguing radio compact source at the center of the Galaxy in 1974 (Brown \\& Lo 1982; Zhao \\etal 1989). We monitored Sgr A* with the VLA during the period of 1990 to 1993. Fluctuations in flux density suggested that the amplitude of variation increased towards short wavelengths and that the rate of outbursts is about three per year (Zhao \\etal 1992; Zhao \\& Goss 1993). Large amplitude fluctuations in the flux density have been observed at millimeter wavelengths (Backer \\& Wright 1993; Tsuboi \\etal 1999). Based on the radio monitoring data obtained with the 3.5 km Green Bank Interferometer at 11 and 3.6 cm, Falcke (1999) reported that at both wavelengths a characteristic time scale of 50 to 200 days is observed while the structure function of 11 cm data suggests a quasi-periodic variation with a period of 57 days. ", "conclusions": "If the 106 day cycle is related to an orbital emitting object around the massive black hole, its distance to the central mass would be 1200$\\times$ R$_g$ (60 AU), where R$_g = 7.5\\times 10^{11}$ cm for 2.5$\\times10^6$ M$_\\odot$. At a distance of 8 kpc, 60 AU corresponds to 8 mas. A 200 mJy compact object separated by 8 mas from Sgr A* is easily detected with the VLBA at wavelengths shorter than 3.6 cm. There is no evidence for a companion source in images of Sgr A* at any wavelength (e.g., Bower \\& Backer 1998). We can also demonstrate the unlikelihood of the specific case of an eccentric binary pair in which a wind from the secondary eclipses the primary as in the case for the pulsar PSR 1259-63 (Johnson \\etal 1992). The opacity of free-free absorption $\\tau_{f-f} \\sim \\lambda^{2.1}$ suggests that the periodic variability due to an orbiting body would be enhanced at longer wavelengths. At $\\lambda \\ge$ 6 cm, Sgr A* would be completely absorbed in the eclipsed phase if $\\tau\\sim 0.5$ as inferred from the fraction of 40\\% in the flux density variation at 1.3 cm. Thus, the periodic variability is more likely intrinsic to Sgr A* and probably occurs within a few hundred R$_g$ of the massive black hole. Quasiperiodic variability in radio flux density can be produced through jets or collimated flows. With a wealth of data observed at wavelengths between radio to X-ray, the microquasar GRS 1915+105 provides an excellent case showing two distinct states ( ``plateau'' and ``flare'') of the accretion process in a stellar mass black hole, (Mirabel \\& Rodriguez 1999; Dhawan \\etal 2000). The current state of Sgr A* is similar to the ``plateau'' state of GRS 1915+105 in terms of flat radio spectrum, compact source size (AU scale) and periodic variability in radio flux density. This state can be contrasted to the ``flare'' state which is characterized by optically thin ejecta feeding large scale jets (500 AU). The radio variability as correlated with the soft X-ray cycle in the ``plateau'' state of GRS 1915+105 suggests that the radio oscillations correspond to the thermal-viscous instability of the accretion disk (Dhawan \\etal 2000). The analog of the radio fluctuations in the two sources suggests that the periodic flux density variations in Sgr A* are also related to an instability of the accretion disk. However, the X-ray luminosity ($\\sim2\\times10^6$ L$_\\odot$) of GRS 1915+105 is 20 times greater than the Eddington limit for a 3 M$_\\odot$ black hole (Mirabel \\& Rodriguez 1999) and the radio jets are produced by the overwhelming radiation pressure. On the other hand, the low X-ray luminosity ($<$ 100 L$_\\odot$, $\\sim$10 orders in magnitude below the Eddington limit for a 2.5$\\times10^6$ M$_\\odot$ object) of Sgr A* indicates that the gravity far exceeds the radiation pressure. Because of the strong gravity and weak radiation pressure, the consequence of the gas dynamics in Sgr A* on the AU scales would be different from GRS 1915+105. In fact, the time variability and the limit on the intrinsic source size ($<$0.5 mas or 100 R$_g$ or 5 AU) of Sgr A* from the 7 mm observations (Lo \\etal 1998 and Bower and Backer 1998) suggest that any variability in the jet occurs in a region where the gravitational field of the black hole dominates. Any collimations of jets or outflows related to the observed radio variability appear to be disrupted within Sgr A* on a scale of $\\sim$5 AU. A model consisting of a jet nozzle ({\\it e.g.} Falcke 1996) seems useful to study if a self-consistent dynamic theory for the disk instability can be constructed. A convection process is now considered in the advection dominated accretion flow (ADAF) that could well provide a reasonable dynamic model for the accretion disk and possible outflows (Narayan \\etal 2000; Quataert \\& Gruzinov 2000; Igumenshchev \\& Abramowicz, 1999; Stone \\etal, 1999). In this theory, hot dense bubbles are produced in the inner part of a low-viscosity disk through convection caused by thermal instability. The most attractive result from the convective-ADAF theory is that quasiperiodic production of convective bubbles has been observed in numerical simulations, although the observed period and the small cycle frequency width have not been predicted in detail from the theory (Igumenshchev \\& Abramowicz, 1999). The current results appear to favor the convective-ADAF model although we can not rule out the possibilities of an orbiting companion object that triggers periodic flares from a jet nozzle." }, "0011/astro-ph0011443_arXiv.txt": { "abstract": "While all but one Gamma-Ray Bursts observed in the X-ray band showed an X-ray afterglow, about 60 per cent of them have not been detected in the optical band. We demonstrate that in many cases this is not due to adverse observing conditions, or delay in performing the observations. We also show that the optically non-detected afterglows are not affected by particularly large Galactic absorbing columns, since its distribution is similar for both the detected and non-detected burst subclasses. We then investigate the hypothesis that the failure of detecting the optical afterglow is due to absorption at the source location. We find that this is a marginally viable interpretation, but only if the X-ray burst and afterglow emission and the possible optical/UV flash do not destroy the dust responsible for absorption in the optical band. If dust is efficiently destroyed, we are led to conclude that bursts with no detected optical afterglow are intrinsically different. Prompt infrared observations are the key to solve this issue. ", "introduction": "The standard external shock synchrotron model (Meszaros \\& Rees 1997; Sari et al. 1998) has been very successful in describing the properties of observed optical afterglows (Wijers et al. 1997; Galama et al. 1998d, Covino et al. 1999). However, for more than half of the afterglows observed in the optical band we did not detect any emission. We will call these Failed Optical Afterglow (FOA) gamma-ray bursts. In all the \\grb~error boxes promptly followed by narrow field X-ray instruments, an X-ray transient (afterglow) has been detected (with the only exception of GRB~990217), while only for $\\sim 40$ per cent of them optical observations have revealed an afterglow at optical wavelengths. This despite the rough similarity of the X-ray afterglow fluxes and the prompt reaction of optical telescopes. Paczynski (1998) ascribes this failed detection to dust extinction pointing out how this interpretation requires the association of bursts with star forming regions. If this is the case, infrared observations should be better suited for the hunt of afterglows, where the extinction plays a reduced role. For this reason, the IR follow-up of GRBs has recently become quite common, and some afterglows (GRB~990705, Masetti et al. 2000; and GRB~000418, Klose et al. 2000a) have been detected in the infrared before being confirmed at optical wavelengths. Yet, we still miss the detection of an IR afterglow without an optical counterpart: such a detection would confirm the role of dust in FOAs. Adding confusion to this picture, observations of extinction in X-ray spectra seem to reveal a very low gas to dust ratio (Vreeswijk et al. 1999, Galama \\& Wijers 2000) which, if common in all GRB environments, would strongly limit the role of dust extinction in the absorption of afterglows in the optical band and, even more, in the NIR. In this paper we show that upper limits derived for FOAs are indeed not consistent with an ``average afterglow'', contrary to what recently claimed by Galama \\& Wijers (2000). We then analyze the properties of detected and non detected optical afterglows in order to check whether the absorption commonly seen in star forming regions can explain the large fraction of FOAs. \\begin{table*} \\begin{tabular}{lllllllllll} \\hline GRB &$F_{X,\\rm NFI}$ &$\\Delta t_x$ &$\\delta_x$ &Ref &$R$ &$\\Delta t_R$ &$\\delta_R$ &Ref &$z$ &Ref \\\\ &$10^{-13}$cgs & h & & & &h & & & & \\\\ \\hline \\hline 970228.12362 &28$\\pm$4 &8 &1.32 &Co97 &21.5$\\pm$0.3 &16.5 &1.73$\\pm$0.12 &Ma98,Ga00 &0.695 &Bl98 \\\\ 970508.904 &7$\\pm$0.7 &6 &1.1 &Am98 &19.77$\\pm$0.1 &52 &1.2$^a$ &Pe98,Ga98a &0.835 &Me97 \\\\ 971214.97272 &4$\\pm$0.4 &6.7 &0.9 &An97 &22.06$\\pm$0.06 &13 &1.20$\\pm$0.02 &Di98 &3.418 &Ku98 \\\\ 980326.88812 &NP &--- &--- &--- &21.25$\\pm$0.03 &11 &2 &Bl99 &--- &--- \\\\ 980329.1559 &7.8$\\pm$0.9 &7 &1.35 &Za98a,b &21.2$\\pm$0.3$^b$ &17 &1.3$\\pm$0.1 &Re99 &--- &--- \\\\ 980425.90915$^c$ &4$\\pm$0.6 &10 &0.2 &Pi00a &15.7$\\pm$0.1 &59.8 &--- &Ga98b &0.0085 &Ti98 \\\\ 980519.51403 &1.4$\\pm$0.3 &9.7 &1.8 &Ni99 &20.4$\\pm$0.1 &15.5 &2.05$\\pm$0.07 &Ha99 & & \\\\ 980613.20215 &1.1$\\pm$0.3 &9 &0.8 &Co99a &22.9$\\pm$0.2 &16.3 &1 &Hj98,Dj98a &1.0964 &Dj98b \\\\ 990123.40780 &110 &5.8 &1.35 &He99a &18.26$\\pm$0.04$^d$ &3.8 &1.12$\\pm$0.03 &Od99,Ga99 &1.6004 &Ku99 \\\\ 990510.36743 &14.7$\\pm$1.8 &8 &1.4 &Ku00 &17.54$\\pm$0.02 &3.5 &0.82$\\pm$0.02 &Hr99 &1.619 &Vr99a \\\\ 990705.66765 &1.9 &11 &1.6 &Am00 &18.7$\\pm$0.05$^e$ &5.5 &1.68$\\pm$0.10 &Ma00 &--- &--- \\\\ 990712.69655 &NP &--- &--- &--- &19.4$\\pm$0.1 &4.16 &0.97$\\pm$0.02 &Sa00a &0.4331 &Vr00 \\\\ 001011.66308 &NP &--- &--- &--- &20.6$\\pm$0.1 &8.4 &1.4 &Go00 &--- &--- \\\\ \\hline 980703.182468 &7.5 &22 &1.3$\\pm$0.25 &Ga98c &21.00$\\pm$0.09 &22.6 &1.39$\\pm$0.3 &Ca99a &0.9662 &Dj98c \\\\ 990308.21883 &--- &--- &--- &--- &18.14$\\pm$0.05 &3.34 &1.2$\\pm$0.1 &Sc99 &--- &--- \\\\ 991208.192269 &--- &--- &--- &--- &18.7$\\pm$0.1 &49.9 &2.15 &Je99a &0.7055 &Di99 \\\\ 991216.671544 &1240$\\pm$40 &4.03 &1.64 &Ta99 &18.49$\\pm$0.05 &10.8 &1.22$\\pm$0.04$^f$ &Ha00 &1.02 &Dj99 \\\\ 000131.62446 &--- &--- &--- &--- &23.26$\\pm$0.04 &84.3 &2.25$\\pm$0.19 &An00 &4.50 &An00 \\\\ 000301.41084 &--- &--- &--- &--- &20.42$\\pm$0.06 &36.5 &1.18$\\pm$0.14 &Sa00b &2.0335 &Ca00a \\\\ 000418.41921 &--- &--- &--- &--- &21.63$\\pm$0.04 &59.3 &0.86$\\pm$0.06 &Kl00 &1.1854 &Bl00 \\\\ 000630.02145 &--- &--- &--- &--- &23.04$\\pm$0.08 &21.6 &1.1$\\pm$0.3 &Je00 &--- &--- \\\\ 000911.30237 &--- &--- &--- &--- &20.26$\\pm$0.17 &34.3 &1.5$\\pm$0.14 &Pr00a,La00 &--- &--- \\\\ 000926.99274 &2.1$\\pm$0.6 &54.2 &4.3$\\pm$1.0 &Pi00b &19.37$\\pm$0.02 &20.7 &1.36$\\pm$0.11 &Sa00c,Fy00a &2.066 &Fy00b\\\\ 001007.20749 & --- &--- &--- &--- &20.3 &83 &0$^g$ &Ca00b,Pr00b &--- &--- \\\\ \\hline \\hline \\end{tabular} \\begin{flushleft} Notes: $\\delta_\\nu$ is defined by $F_\\nu (t) \\propto t^{-\\delta_\\nu}$. X-ray fluxes in the 2--10 keV band. NP=repointing of \\sax~not possible. $^a$: for $t>2$ days; earliest detection at 3.1 hours: $R=21.1\\pm0.1$. $^b$: $R$ mag derived from $I=20.8\\pm0.3$; $R$=23.6$\\pm$0.2 after 20 hours. $^c$: = SN 1998bw, not used in the analysis. $^d$: Converted from Gunn $r$-mag. $^e$: $R$ mag derived from $H=16.57\\pm0.05$. $^f$: for t$\\le$1.2 days. $^g$: for t$\\le$3.5 days, $\\delta_R\\sim1.4$ after.\\\\ Am98: Amati et al., 1998; Am00: Amati et al., 2000 An97: Antonelli et al., 1997; An00: Andersen et al., 2000; Bl98: Bloom et al., 1998; Bl99: Bloom et al., 1999; Bl00: Bloom et al., 2000; Ca99a: Castro-Tirado et al., 1999a; Ca00a: Castro et al., 2000a; Ca00b: Castro et al., 2000b; Co97: Costa et al., 1997; Co99a: Costa et al., 1999a; Di98: Diercks et al., 1998; Dj98a: Djorgovski et al., 1998a; Dj98b: Djorgovski et al., 1998b; Dj98c: Djorgovski et al., 1998c; Dj99: Djorgovski et al., 1999; Do99: Dodonov et al., 1999; Fy00a: Fynbo et al., 2000a; Fy00b: Fynbo et al., 2000b; Ga98a: Galama et al., 1998a; Ga98b: Galama et al., 1998b; Ga98c: Galama et al., 1998c; Ga99: Galama et al., 1999; Ga00: Galama et al., 2000; Go00: Gorosabel et al., 2000; Ha99: Halpern et al., 1999; He99a: Heise et al., 1999; Hj98: Hjorth et al., 1998; Hr99: Harrison et al., 1999; Je99a: Jensen et al., 1999a; Je00: Jensen et al., 2000; Kl00: Klose et al., 2000a; Ku98: Kulkarni et al., 1998; Ku99: Kulkarni et al., 1999; Ku00: Kuulkers et al., 2000; La00: Lazzati et al., 2000; Ma98: Masetti et al., 1998; Ma00: Masetti et al., 2000; Me97: Metzger et al., 1997; Ni99a: Nicastro et al., 1999a; Od99: Odewahn et al., 1999; Pe98: Pedersen et al., 1998; Pi00a: Pian et al., 2000; Pi00b: Piro et al., 2000; Pr00a: Price et al., 2000a; Pr00b: Price et al., 2000b; Re99: Reichart et al., 1999; Sa00a: Sahu et al., 2000; Sa00b: Sagar et al., 2000; Sa00c: Sagar et al., 2001; Sc99: Schaefer et al., 1999; Ta99: Takeshima et al., 1999; Ti98: Tinney et al., 1998; Vr99a: Vreeswijk et al., 1999a; Vr00: Vreeswijk et al., 2001; Za98a: In't Zand et al., 1998a; Za98b: In't Zand et al., 1998b; \\end{flushleft} \\caption{{Properties of the bursts with associated optical transient. The first 13 bursts have been observed by the \\sax-GRBM/WFC, while the remaining bursts (below the horizontal line) have been discovered by other instruments (see text).} \\label{tab:uno}} \\end{table*} \\begin{table*} \\begin{tabular}{llllllll} \\hline GRB &$F_{X,\\rm NFI}$ &$\\Delta t_x$ &$\\delta_x$ &Ref &$R$ &$\\Delta t_R$ &Ref \\\\ &$10^{-13}$cgs & h & & & &h & \\\\ \\hline \\hline 970402.930 &2.2$\\pm$0.6 &8 &1.6 &Ni98 &21 &18.5 &Gr97a \\\\ 971227.34938 &2.6$\\pm$0.6 &14 &1.12 &An99a &22.8 &21.3 &Gr97b \\\\ 981226.40793 &5$\\pm$1 &11 &1.3 &Fr00 &23 &10. &Li99 \\\\ 990217.22462 &$<1$ &6 &$>$1.6 &Pi99a &23.5 &19. &Pa99a \\\\ 990627.20894 &3.5 &8 &--- &Ni99b &21. &23. &Ro99 \\\\ 990704.7294 &4.4$\\pm$0.3 &8 &--- &Fe99 &22.5 &4.6 &Je99b \\\\ 990806.60286 &5.5$\\pm$1.5 &7.8 &--- &Fr99 &22. &3.8 &Vr99c \\\\ 990907.7319 &15$\\pm$5 &11 &--- &Pi99b &22.9$^a$ &24.9 &Pa99b \\\\ 990908.00125 &NP &--- &--- &--- &20.$^b$ &11.5 &Ax99 \\\\ 991014.9115 &3.5$\\pm$0.5 &13 &$>$0.4 &Za00 &22.6 &12.9 &Ug99 \\\\ 991105.69495 &NP &--- &--- &--- &23.5 &16. &Pa99c \\\\ 991106.4545 &1.25$\\pm$0.3 &8 &--- &An99b &21. &9.1 &Ca99b \\\\ 000210.36396 &4.5 &7.2 &--- &Co99b &23.3 &16. &Go00a,b \\\\ 000214.042 &2.75$\\pm$0.9 &12 &0.6 &An00 &21.$^c$ &32.4 &Rh00 \\\\ 000424.76258 &--- &--- &--- &--- &22.8 &33. &Ug00 \\\\ 000528.36568 &1.7$\\pm$0.3 &8.3 &1 &Ku00 &23.3 &18 &Pa00a \\\\ 000529.3361 &2.8$\\pm$0.7 &7.5 &--- &Fe00 &22.3 &47 &Pa00b \\\\ 000615.2625 &--- &10 &--- &BS00 &21.5 &4.2 &St00 \\\\ 000620.2317 &--- &--- &--- &--- &19.8 &5.7 &Go00c \\\\ \\hline 990520.08539 &--- &--- &--- &--- &21.7$^d$&19.5 &Ma99 \\\\ 991217.17496 &--- &--- &--- &--- &22. &11. &Mo99 \\\\ 000416.6062 &--- &--- &--- &--- &20.7 &50.3 &Pr00c \\\\ \\hline \\hline \\end{tabular} \\\\ \\begin{flushleft} Notes:NP=Repointing of \\sax~not possible; $^a$: $R$ mag derived from V$>23.2$. $^b$: $R$ mag derived from $V>20.3$. $^c$: $R$ mag derived from $K>18.15$. $^d$: $R$ mag derived from $V>22$. \\\\ An99a: Antonelli et al., 1999a; An99b: Antonelli et al., 1999b; An00: Antonelli et al., 2000; Ax99: Axwlrod et al., 1999; BS00: \\sax~mail \\# 00/18 = GCN Circ. \\# 707; Ca99b: Castro-Tirado et al., 1999b; Co99b: Costa et al., 1999b; Fe99: Feroci et al., 1999; Fr99: Frontera et al., 1999; Fr00: Frontera et al., 2000; Go00a: Gorosabel et al., 2000a; Go00b: Gorosabel et al., 2000b; Go00c: Gorosabel et al., 2000c; Gr97a: Groot et al., 1997a; Gr97b: Groot et al., 1997b; Li99: Lindgren et al., 1999; Je99b: Jensen et al., 1999b; Ma99: Masetti et al., 1999; Mo99: Mohan et al., 1999; Ni98: Nicastro et al., 1998; Ni99b: Nicastro et al., 1998; Pa99a: Palazzi et al., 1999a; Pa99b: Palazzi et al., 1999b; Pa99c: Palazzi et al., 1999c; Pa00a: Palazzi et al., 2000a; Pa00b: Palazzi et al., 2000b; Pi99a: Piro et al., 1999a; Pi99b: Piro et al., 1999b; Pr00c: Price et al., 2000; Rh00: Rhoads eta l., 2000; Ro99: Rol et al., 1999; St00: Stanek et al., 2000; Ug99: Uglesich et al., 1999; Ug00: Uglesich et al., 2000; Vr99c: Vreeswijk et al., 1999c; Za00: in't Zand et al., 2000. \\end{flushleft} \\caption{{Properties of the bursts with \\sax-WFC detection but without associated optical transient. The last three bursts below the horizontal line refer to the $\\gamma$-ray poor GRBs (or X-ray transients) detected by \\sax.} \\label{tab:due}} \\end{table*} ", "conclusions": "By analyzing the properties of detected optical and X-ray afterglows and the upper limits for failed detections, we show that the subset of bursts without optical afterglow (FOAs) defines a different family. This conclusion relies on several assumptions, like the homogeneity of the \\sax~and non~\\sax~detected afterglows, imposed by the paucity of the sample. Should some of these turn out to be wrong, the conclusion would become statistically less stringent (see \\S2 for further details). We have investigated if this can be due to dust extinction of optical radiation in a molecular cloud. We find that this hypothesis can only marginally account for the large fraction of FOAs, and therefore we cannot exclude the possibility that FOAs are intrinsically less luminous in the optical/UV band with respect to the detected ones, and with respect to their own X-ray luminosity. Consider also that we have been very conservative in our procedure, because our results are based on considering {\\it upper limits} on the optical flux, and {\\it peak} absorption columns expected in giant molecular clouds. The latter assumptions may well be too conservative, if the dust is bound to evaporate when illuminated and heated by the powerful optical/UV flash of the gamma-ray burst (Waxman \\& Draine 2000) and by its X-ray radiation (Fruchter et al. 2000). This dust sublimation is suggested for a sample of burst afterglows (Vreeswijk et al. 1999c, Galama \\& Wijers 2000), in which a very large hydrogen column density $N_{\\rm H} \\gsim 10^{22}$ cm$^{-2}$, as estimated by X-ray data, is associated with almost no optical extinction. The results can be understood only in terms of a dust to gas ratio $\\sim 100$ times smaller than the Galactic average value. In turns, such low values of the dust to gas ratio can be explained only if the dust has been completely sublimated in the surroundings of the burst. Indeed the theoretical models mentioned above predict that dust can be destroyed by the burst emission out to a radius comparable to the dimension of a typical molecular cloud (up to a few tens of parsecs). If this is the case, the material responsible for absorption in FOAs is not the overdense cocoon surrounding the star forming region, but the cloud as a whole (or even less), and the discrepancy between the observed and measured value (see dash-dotted line in Fig.~\\ref{fig:mix}) becomes extremely compelling. An interesting way to assess whether dust is playing any role in FOAs is to perform near infrared (NIR) follow-up of their $\\gamma$ or X-ray error boxes. For instance, in the $K$ filter, absorption is greatly reduced, so that only a very small fraction (less than 10 per cent) of afterglows should show more than 1 magnitude of absorption, in any of the adopted cloud models. This is therefore a crucial test to understand whether FOAs are due to dust absorption (less severe in the near infrared) or to an intrinsic difference in the emitted spectrum (that should be more severe in the NIR). Some FOAs have been indeed looked for in the NIR band, but the observations are still very sparse and we lack any statistics to draw any meaningful conclusion. NIR observations are thus strongly recommended as the key test for the dust extinction hypothesis, especially after the launch of HETE II, which will rapidly distribute accurate enough locations of bursts to be promptly followed by ground based telescopes. A more homogeneous dataset, though, will have to await the launch of the Swift satellite, foreseen in 2003. The systematic follow-up with the on-board optical telescope will provide a multiband spectroscopic database of the first hours of optical afterglows. Data of even higher quality could be achieved if IR robotic telescopes (such as the one proposed by the consortium of Brera, Rome and Catania Observatories, called REM, for Rapid Eye Mount), will be in operation to complement Swift observations from the ground. A possibility to increase the absorption in the observed $R$ band without invoking particularly dense molecular clouds is by allowing for a higher redshift of the bursts. This would make the afterglow undetectable, especially if the redshift of the burst is particularly high ($z\\gsim 4$), so that the redshifted Lyman $\\alpha$ break falls in the $R$ filter. In this case, again, near infrared ($JHK$) observations should be unaffected by absorption and the optical transient easily detectable. Estimates of the fraction of high redshift GRBs (see, e.g., Porciani \\& Madau 2001) predict however a very small fraction of bursts (up to few per cent) at $z>3$, if the GRB and star-formation rate are related. A more interesting way-out is that the property of clouds at high redshift are different from those of our Galaxy (see Ramirez-Ruiz Trentham \\& Blain 2001), or that the dust extinction curve changes its shape with redshift. In conclusion, we have found that dust absorption due to a cloud with properties similar to Galactic clouds is not a completely satisfying explanation for bursts without a detected optical afterglow, and we cannot rule out the possibility that they are due to an intrinsic larger dispersion of optical fluxes with respect to the dispersion of the X-ray fluxes (see also B\\\"oer \\& Gendre 2000). This, in turn, opens some exciting observational perspectives aiming to disclose the nature of the burst progenitor: if bursts are indeed associated with the final stages of stellar evolution and a supernova-like event is associated to all bursts, then the search for supernova signatures should be easier for bursts with an optical faint afterglow, for which the SN lightcurve would not be polluted by the flux of the afterglow. If we assume SN1998bw (Galama et al. 1998c) as a template supernova lightcurve, the expected magnitudes at maximum should be roughly $I=24$ and $R=25$, easily detectable with a signal to noise ratio of $\\sim 10$ with an exposure time of only $\\sim 10$~min with an 8 meter class telescope." }, "0011/astro-ph0011505_arXiv.txt": { "abstract": "HST V and I-band observations show that the gravitational lens B1359+154 consists of six images of a {\\it single} $z_s=3.235$ radio source and its star-forming host galaxy, produced by a compact group of galaxies at $z_l \\simeq 1$. VLBA observations at 1.7 GHz strongly support this conclusion, showing six compact cores with similar low-frequency radio spectra. B1359+154 is the first example of galaxy-scale gravitational lensing in which more than four images are observed of the same background source. The configuration is due to the unique lensing mass distribution: three primary lens galaxies lying on the vertices of a triangle separated by $0\\farcs7 \\simeq 4 h^{-1}$ kpc, inside the $1\\farcs7$ diameter Einstein ring defined by the radio images. The gravitational potential has additional extrema within this triangle, creating a pair of central images that supplement the ``standard'' four-image geometry of the outer components. Simple mass models consisting of three lens galaxies constrained by HST and VLBA astrometry naturally reproduce the observed image positions but must be finely-tuned to fit the flux densities. ", "introduction": "The vast majority of the $\\sim 60$ known arcsecond-scale gravitational lens systems consist of two or four detectable images, consistent with the generic lensing properties of smooth, isolated and centrally steep mass distributions (e.g.\\ Blandford \\& Kochanek 1987). There are very few cases in which a ``non-standard'' number of images may have been observed.\\footnote{B1933+503 (Sykes et al.\\ 1998) and B1938+666 (King et al.\\ 1997) contain ten and six lensed radio components, respectively, but this is due to the imaging of a multi-component source rather than any exotic properties of the lensing potential.} APM08279+5255 (Ibata et al. 1999) and MG1131+0456 (Chen \\& Hewitt 1993) each contain a central component that could be an additional lensed image created by a sufficiently large galaxy core or shallow mass profile (Narasimha, Subramanian \\& Chitre 1986; Rusin \\& Ma 2000). Alternatively, APM08279+5255 may be a special class of imaging produced by an edge-on disk (Keeton \\& Kochanek 1998), while the central component in MG1131+0456 could be weak AGN emission associated with the lensing galaxy. In addition, MG2016+112 (Lawrence et al. 1984; Garrett et al. 1996) exhibits a rather complicated image morphology consisting of four primary components, one of which has three subcomponents in VLBI maps. This may be a compound lens produced by two galaxies at different redshifts (Nair \\& Garrett 1997). Benitez et al.\\ (1999) suggest an alternative model, but it might not fully explain the VLBI observations. Complex interplay between mass distributions can lead to lens systems with more than four images of a source. For example, ellipsoidal mass distributions perturbed by shear fields may produce configurations with six or eight images arranged about the tangential critical curve (Keeton, Mao \\& Witt 2000b). However, this requires that the relative magnitudes and orientations of the internal and external shear axes be finely tuned, and the resulting cross-sections are quite small. Compound mass distributions have been shown to be far more efficient at producing a variety of complex geometries in which five or more images may be formed over a range of radii (Kochanek \\& Apostolakis 1988). This has been observed in the cluster-lensing case, where substructure in the gravitational potential created by the resident galaxies can qualitatively alter the lensing properties that one would expect for a smooth halo mass distribution (e.g. Natarajan et al. 1998; Meneghetti et al. 2000). A dramatic example of this is CL0024+1654, which exhibits eight images of a single blue background galaxy (Kassiola, Kovner \\& Fort 1992; Wallington, Kochanek \\& Koo 1995; Colley, Tyson \\& Turner 1996; Tyson, Kochanski \\& Dell'Antonio 1998). Early-type galaxies preferentially participate in lensing (Kochanek et al.\\ 2000a), so a large fraction of lens galaxies should be members of small groups and clusters due to the morphology-density relation (Dressler 1980; see also Keeton, Christlein \\& Zabludoff 2000a). Indeed, many lens systems are known to be perturbed by nearby galaxies (e.g.\\ B2319+051; Rusin et al.\\ 2000b), groups (e.g.\\ PG1115+080; Schechter et al.\\ 1997), or clusters (e.g.\\ QSO 0957+561, Fischer et al.\\ 1997; RXJ0921+4529, Mu\\~noz et al.\\ 2000). A handful of systems are lensed by more than one primary galaxy (e.g. B1127+385, Koopmans et al.\\ 1999; B1608+656, Koopmans \\& Fassnacht 1999), but in these cases standard image geometries are produced despite merged caustics. Several additional lenses are observed to have faint satellite galaxies as companions inside or near the Einstein radius (e.g.\\ MG0414+0534, Schechter \\& Moore 1993; B1030+074, Xanthopoulos et al.\\ 1998; B1152+199, Rusin et al.\\ in preparation). While the likelihood of finding nearly equal mass galaxies close enough to have merged caustics is predicted to be only $\\sim 1\\%$ (Kochanek \\& Apostolakis 1988), the probability of finding fainter satellites near the primary lens is not small because galaxy luminosity functions diverge for faint galaxies -- all lenses should have faint neighbors, as discussed in the Appendix. While such systems typically produce regions of the source plane in which more than four images can form, the small size of the companion galaxies means that the cross-sections for creating non-standard image geometries are not significant. The gravitational lens system B1359+154 (Myers et al.\\ 1999), discovered in the Cosmic Lens All-Sky Survey (CLASS; e.g.\\ Myers et al.\\ 1995), has been suspected of containing more than four lensed images. Observations with the Very Large Array (VLA) and Multi-Element Radio-Linked Interferometer Network (MERLIN) show a total of six radio components (Myers et al.\\ 1999; Rusin et al.\\ 2000a). Four of these components (A--D) are arranged in a typical quad-lens configuration (maximum image separation of $1\\farcs7$), with two additional components (E and F) residing within the ring defined by the outer images. Preliminary radio spectral studies of B1359+154 at high frequency (Myers et al.\\ 1999) suggested that E had a slightly flatter spectrum than A--D, and therefore that the central components may be core-jet emission associated with a weak AGN in the lensing galaxy or galaxies, as in the case of B2045+265 (Fassnacht et al.\\ 1999). When the spectra were extended down to 5 GHz, however, there appeared to be less disparity among the radio components (Rusin et al.\\ 2000a). Subsequent VLA observations at 15~GHz have failed to decisively confirm that E is flatter than A--D at high-frequency, or detect component F. Spectroscopy with the Keck II telescope determined the source redshift to be $z_s = 3.235$ (Myers et al.\\ 1999). Adaptive optics observations of B1359+154 conducted with the Canada-France Hawaii Telescope (CFHT) in the infrared K$'$ band (Rusin et al.\\ 2000a) detected counterparts to radio components A--D, and discovered three extended emission peaks (K1--K3) bracketing the expected positions of E and F. K1--K3 were identified as three possible lensing galaxies, comprising the core of a compact galaxy group. Evidence of an arc connecting A, B and C was also observed, along with a weaker emission feature associated with component E. The compound deflector system not only explained why attempts to model the outer four components using a single galaxy had failed (Myers et al.\\ 1999), but offered the means of creating additional extrema in the lensing potential. This opened the possibility that at least one of the central components is a lensed image. In this paper we present powerful new evidence from observations with the Hubble Space Telescope (HST) and Very Long Baseline Array (VLBA) that B1359+154 consists of six images of a {\\em single} background source, lensed by a compact group of galaxies at $z_l\\simeq 1$. In \\S2 we present VLBA observations of B1359+154 and investigate the low-frequency radio spectra of the components. In \\S3 we discuss and analyze HST V and I-band observations, which offer compelling evidence for the six-image hypothesis. In \\S4 we use preliminary mass modeling to demonstrate that B1359+154 can be naturally explained as a true six-image lens system. Finally, in \\S5 we discuss the prospects for obtaining improved constraints on the lensing mass distribution of B1359+154, and ultimately using the system to study the structure of small galaxy groups at high redshift. ", "conclusions": "HST and VLBA observations demonstrate that the gravitational lens B1359+154 consists of six images of a single background radio source. This is the first example of a galaxy-scale lens system with more than four images. The unique configuration is produced by the complex mass distribution of the lens, a compact group of three $z_l\\simeq 1$ galaxies lying on the vertices of a triangle separated by $0\\farcs7 \\simeq 4 h^{-1}$kpc, inside the $1\\farcs7$ diameter Einstein ring defined by the radio components. The outer images (A--D) have the morphology of a standard four-image lens, but the triangle of galaxies produces additional extrema in the gravitational potential at which two central images (E and F) form. Simple lens models consisting of three deflectors constrained by HST galaxy coordinates and VLBA radio component data naturally produce six images at the observed positions, but require finely-tuned ellipticities to account for the flux density ratios. While it would be premature to claim that we have accurately represented the lensing mass distribution ($\\chi^2_{tot}$/NDF $\\geq 10$ for all models), our initial modeling efforts demonstrate that B1359+154 can be explained as a true six-image lens system produced by three lensing galaxies. The consistency between the model-predicted deflector mass ratios and the observed luminosity ratios of the lens galaxies adds credibility to our results. The three primary lens galaxies appear to constitute the core of a compact group. The most concentrated groups with small numbers of galaxies are the Hickson groups, a quarter of which have 3 members (Hickson et al. 1992). Separated by only $\\simeq 4 h^{-1}$ kpc, the three galaxies are close enough so that they might be expected to be consumed in a merger on a time scale of $\\sim 0.1$ $H_o^{-1}$ (Barnes 1985). Thus the incidence of high multiplicity arcsecond-scale lens systems may allow for an estimate of the merger rate of massive galaxies at $z \\simeq 0.5 -1.5$. The relative simplicity of B1359+154 compared to the lenses found in the cores of rich clusters should make this system an excellent tool for studying galaxy halos in dense environments, and the relationship between the galaxies and their parent halos. One issue worthy of investigation is whether the galaxy mass distributions are truncated on the scale of inter-galaxy separation. The improved fits of the 3PJS and 3PJE/FIX models relative to the 3SIS and 3SIE/FIX models offer tentative arguments to this effect. Because predicted time delays in the PJ models are significantly higher than those in the SIE models, measured delays may be used to discriminate between the profiles. A second interesting issue is the placement of the group halo relative to the lensing galaxies. The image positions and flux densities can be reproduced by three mass distributions associated with the observed galaxies, so there is little need to add a fourth independent mass distribution inside the Einstein ring based on modeling arguments. A fourth deflector above the critical density for multiple imaging would place an additional radial caustic onto the source plane and disturb the delicate balance needed to account for the data without creating additional images. Thus if this halo exists within the Einstein ring, it is likely to be either subcritical (large core radius) or centered on one of the galaxies. However, the similar model-predicted mass-to-light ratios of the galaxies argue against the latter possibility. The large external shear fields ($\\gamma \\gtorder 0.15$) required by all of our models suggests that a group halo may be displaced from the primary lenses. It is interesting to note that the orientation of the shear axis is remarkably constant, running nearly east to west regardless of the deflector model employed. The six lensed components allow for more complicated models than we have investigated here, but the development of a truly robust model of the lensing mass distribution requires the acquisition of additional constraints. The detection of lensed extended emission from the quasar host galaxy in both the HST optical and CFHT infrared images is particularly promising. Deep, high-resolution HST imaging using either WFPC2 or NICMOS could investigate the properties of the Einstein ring in detail and provide vital new constraints on the potential structure (see Kochanek et al.\\ 2000b). Furthermore, if subcomponents of D, E and F can be detected using deep global VLBI imaging, the relative orientations of the radio substructure could place additional constraints on the local magnification matrices. Finally, the measurement of differential time delays would also offer important constraints on the mass model, as these directly probe the lensing potential at the image positions. One might expect the source to be variable, based on the compactness of the radio components, though there is no evidence for significant variability at this point. Radio snapshots of B1359+154 should be routinely obtained to determine whether the variability is large enough to undertake a monitoring program using current instruments. Otherwise, the enhanced VLA may be able to determine time delays from the component light curves even if the source variability is only at the few percent level. It seems unlikely that the lens is a chance projection of three galaxies (an {\\it a posteriori} probability of $\\sim 10^{-3}$ in a sample of 60 lenses), and their statistically identical colors further argues against this possibility. While the Next Generation Space telescope could directly determine the redshifts of all three galaxies, current telescopes can only measure the average redshift of the lenses unless they have narrow emission lines that are not apparent in the Myers et al. (1999) spectrum. Two possibilities are to search for Mg{\\sc~II} metal absorption lines in the source spectrum or HI absorption features in the radio continuum once the average redshift is known. For $z_l\\simeq1$ the Mg{\\sc~II} absorption features would lie in the V-band where the source flux is 23~V~mag, so the observation is difficult. Many groups have significant HI masses (Hickson 1997) and the radio flux density of $\\simeq 90$~mJy at 0.7~GHz (based on the 1.7 GHz VLBA flux density and a spectral index of $\\alpha \\simeq 0.3$ at low frequency) may be high enough to search for absorption features. X-ray observations are less promising because very deep observations would be required to detect the group (a bright $10^{42}$~erg~s$^{-1}$ group at $z_l=1$ has an X-ray flux of only $4\\times 10^{-16}$~ergs~cm$^{-2}$~s$^{-1}$) and the high resolution of the Chandra Observatory would be needed to distinguish emission from the group and the lensed sources. \\noindent Acknowledgements: The authors deeply thank the director of STScI, Steve Beckwith, for making these observations possible. We also thank Lars Hernquist and Ann Zabludoff for discussions about groups and the implications of this lens. DR acknowledges support from the Zaccheus Daniel Foundation. Support for the CASTLES project was provided by NASA through grant numbers GO-8175, GO-8268 and GO-8804 from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc. CSK, EEF, JL, BAM and JAM were also supported by the Smithsonian Institution. CSK and CRK were also supported by the NASA Astrophysics Theory Program grant NAG5-4062. HWR is also supported by a Fellowship from the Alfred P. Sloan Foundation. \\appendix" }, "0011/hep-ph0011030_arXiv.txt": { "abstract": "If neutrino oscillation plays a role in explaining the atmospheric neutrino deficit, then the same phenomenon would necessarily affect also the dark matter indirect-detection signal which consists in a muon-neutrino flux produced by neutralino annihilation in the Earth core. In this paper we investigate to which extent the upgoing-muon signal originated by neutralinos captured inside the Earth would be affected by the presence of $\\nu_\\mu \\rightarrow \\nu_\\tau$ oscillation. ", "introduction": "Among the different techniques which have been proposed to search for dark matter particles \\cite{nota0}, detection of a neutrino flux by means of neutrino telescopes represents certainly an interesting tool. Although direct detection \\cite{ICTP,JKG,old} at present appears to be somewhat more sensitive to neutralino dark matter \\cite{probing,further,notadir}, nevertheless all the different possibilities are worth being explored. In this paper we discuss the flux of upgoing muons which are a consequence of $\\nu_\\mu$'s produced inside the Earth, with special emphasis on the role which is played in this kind of searches by the possible presence of neutrino oscillation. ", "conclusions": "In this paper we have discussed to which extent neutrino oscillation can affect the up--going muon signal from neutralino annihilation in the Earth. While the experimental upper limit is, at present, practically not affected by neutrino oscillation \\cite{MACRO}, the theoretical predictions are reduced in the presence of oscillation. By adopting the neutrino oscillation parameters deduced {}from the fits on the atmospheric neutrino data \\cite{FGGV,SK_OSC_nos}, the effect is always larger for lighter neutralinos. For $\\nu_\\mu \\rightarrow \\nu_\\tau$ the reduction is between 0.5 and 0.8 for $m_\\chi \\lsim 100$ GeV and less than about 20\\% for $m_\\chi \\gsim 200$ GeV." }, "0011/hep-ph0011176_arXiv.txt": { "abstract": "We review the phenomenology of UHE neutrino detection. The motivations for looking for such neutrinos, stemming from observational evidence and from the potential for new physics discoveries are enumerated, and their expected sources and fluxes are given. Cross-sections with nucleons at energies all the way upto $10^{20}$ eV and the attenuation of fluxes in the Earth, both of which are physics issues important to their detection, are discussed. Finally, sample event-rates for extant and planned Water/Ice Cerenkov detectors are provided. ", "introduction": "It is now widely beleived that the recent Super Kamiokande (SK) result {\\cite{fuk,sobel}} of an anomaly in the flavour ratios and zenith angle dependance of the atmospheric neutrino flux provides the firmest signal yet of physics beyond the Standard Model (SM)\\footnote{The first hints of this anomaly were provided by atmospheric neutrino data from the IMB detector, \\cite{IMB}} . The significance of this can be gauged by the fact that a signal for such physics has been sought for in all the varied and intensive experimental probes that the many sectors of the SM have been subjected to over more than twenty-five years. That this signal comes from the neutrino sector of the theory perhaps assumes increased significance when considered in conjuntion with the fact that two other existing experimental anomalies which persist, the solar deficit \\cite{solar} and the LSND result \\cite{lsnd} are also within the neutrino sector. The detailed interpretation of these anomalies and their relation with each other is still a matter of intense and ongoing theoretical and experimental activity. However, neutrino masses, mixings and consequent oscillations have provided the simplest framework for understanding the experimental results \\cite{smir}. What may be said with a reasonable degree of conviction is that once their interpretation is clear, a hitherto unprecedented window into the nature of the physical theory beyond the SM will probably have been opened. It is within this context, perhaps, that one should view the theoretical and experimental efforts that are ongoing in the area of ultra high energy (UHE) neutrino physics (for reviews see \\cite {gqrs,h1,r98}). If the physics excitement that we have uncovered in our study of low-energy neutrinos is an indicator, then the study of UHE neutinos should pay rich dividends. In view of the fact that data collection and further upgradation is ongoing for the first generation of UHE neutrino detectors ( AMANDA \\cite{aman} and BAIKAL \\cite{bai}) and that planning, design, testing and deployment is underway for some others (AUGER \\cite {aug}, NESTOR \\cite{nest}, ICECUBE \\cite{ice}, ANTARES \\cite{ant}, RICE \\cite{rice}), and NEMO \\cite{nemo}) definitive results of these efforts are still in the future. However, besides the other specific reasons from astrophysics and cosmic ray physics (some of which we discuss below) which provide motivation for such experiments, we note that the energy range covered by these experiments ($E_{\\nu} \\simeq 1$ Tev to $10^9$ TeV or higher) offers an unprecedented opportunity for particle physics at energies significantly beyond the scope of terrestial accelerators. Although progress is expected to be slow, the potential for serendipitous discovery is undoubtedly high. In what follows, we give specific reasons why the search for UHE neutrinos can be expected to yeild positive results. Possible sources and their fluxes are discussed. Salient features pertinent to the phenomenology of detection are then outlined, and sample rates provided. ", "conclusions": "We have made an attempt to review the essential motivations and phenomenological issues related to the interactions and detection of UHE neutrinos and provided expected event rates for water/ice Cerenkov detectors. Although present detector capabilities and sizes do not allow them to see above atmospheric backgrounds, with upgradations and several new large-scale experiments underway, the detection of UHE neutrinos from astrophysical sources may be very likely in the near future. Besides the potential for serendipitous discovery, the unmatched (terrestially) energies may lead to new particle physics discoveries. In addition, these experiments may help clarify the astrophysical mechanism responsible for the ultra-relativistic acceleration of matter in these sources and answer important questions in CR physics, and, finally, provide sensitive tests of gravitational couplings and oscillations. For those working in UHE neutrino physics and astronomy, the anticipation is palpable." }, "0011/astro-ph0011263_arXiv.txt": { "abstract": "Here we report the detection, in H$\\alpha$ emission, of a radial corrugation in the velocity field of the spiral galaxy NGC 5427. The central velocity of the H$\\alpha$ line displays coherent, wavy-like variations in the vicinity of the spiral arms. The spectra along three different arm segments show that the maximum amplitude of the sinusoidal line variations are displaced some 500 pc from the central part of the spiral arms. The peak blueshifted velocities appear some 500 pc upstream the arm, whereas the peak redshifted velocities are located some 500 pc downstream the arm. This kinematical behavior is similar to the one expected in a galactic bore generated by the interaction of a spiral density wave with a thick gaseous disk, as recently modeled by Martos \\& Cox (1998). ", "introduction": "It has been known for decades that the gaseous disk of our Galaxy displays coherent vertical distortions (Kerr 1957; Gum, Kerr \\& Westerhout 1960). In particular, for locations inside the solar circle, a wavy structure has been observed in the vertical distribution of several interstellar tracers, including the atomic and ionized components (Lockman 1977; Sanders, Solomon \\& Scoville 1984; Spicker \\& Feitzinger 1986; Malhotra 1995), and young stellar objects (\\eg Dixon 1967; Alfaro \\etal 1992a,b; Berdnikov \\& Efremov 1993). These undulations, distributed above and below the mean Galactic plane and in both the azimuthal and radial directions, have been termed as ``corrugations'', and they seem to be common features appearing in disk galaxies. They have also been detected in both the gaseous and stellar disks of other spiral galaxies such as M31 (Arp 1964), NGC 4244, and NGC 5023 (Florido \\etal 1991). Their origin could be ascribed to either well localized features, such as spiral arms and collisions with high-velocity clouds (HVC), or large scale perturbations, such as gravitational interactions. For instance, some authors have explored the possibility of undulations along spiral arms, induced by magneto-gravitational instabilities (\\eg Nelson 1985; G\\'omez de Castro \\& Pudritz 1992; Kim \\etal 1997; Franco \\etal 2000). In this case, the arms are corrugated by the undular mode (Parker 1966) of the instability, and the structuring occurs $only$ in the azimuthal direction (a possible association to the radial patterns could be made if one assumes that these are the projection of the azimuthal corrugations on the radial direction). Another localized origin could be ascribed to collisions of HVC with the disk of a galaxy: HVC impacts could cause the midplane disk to oscillate in a mode of wavy patterns as well (Franco \\etal 1988; Santill\\'an \\etal 1999). In the large scale scenario, the undulations represent the natural response of the disk, which behaves like a loose drumhead, to the perturbations induced either by some companion galaxies or other galactic subsystems (Weinberg 1991; Edelsohn \\& Elmegreen 1997). One obvious inference is that these corrugations should be associated with equally undulated coherent motions, and models addressing the vertical displacements also show wave-like $z$-velocity fields with amplitudes of the order of tens of km~s$^{-1}$ (\\eg Nelson 1976, 1985; Franco \\etal 1989, 2000; Edelsohn \\& Elmegreen 1997; Kim \\etal 1997; Santill\\'an \\etal 2000). Despite this expectation, however, the data about actual velocity features associated with spatial corrugations are scarce. The observational evidence for small-scale distortions in the vertical velocity fields of disk galaxies comes from the so-called {\\sl rolling motions} in our Galaxy (observed in the $b - V$ plots of the HI galactic distribution; Yuan \\& Wallace 1973; Feitzinger \\& Spicker 1985), the velocity field of the ionized gas in M 51 (Goad, De Veny \\& Goad 1979), rotation curves of some spirals (Pismis 1965; Rubin, Ford \\& Thonnard 1980), and optical and radio isovelocity maps of some field galaxies (Bosma 1978; Pismis 1986). In particular, the study of the velocity field of the ionized gas in M 51 performed by Goad, De Veny \\& Goad (1979) suggests that part of the observed nonrotational velocities could be due to motions perpendicular to the plane of the disk. Pismis (1965, 1986), on the other hand, has pointed out the relevance of analyzing the location of the velocity undulations in the rotation curve with respect to the spiral structure of the host galaxy. Clearly, all this kinematic information may be important in our understanding of corrugations. In this paper we report the detection of a corrugated velocity field in the ionized gas of the nearly face-on spiral NGC 5427. The central velocity of the H$\\alpha$ emission line shows velocity shifts with similar behavior when the gas passes through the spiral arms. This behavior can be understood as due to a hydraulic jump, or bore, generated by the interaction of the spiral density wave with the gas of a thick gaseous disk, as discussed by Martos \\& Cox (1998, hereafter MC) and Martos \\etal (1999). The detection of this velocity pattern in NGC 5427 does not answer the current questions about the morphogenesis of corrugations, but represents a powerful tool to explore further the theoretical and observational aspects of the problem. The paper has been organized in three sections, the first one being this introduction; section 2 is devoted to the description of the observations and reduction procedure and, finally, in section 3, we discuss the results and outline the main conclusions. ", "conclusions": "" }, "0011/astro-ph0011055_arXiv.txt": { "abstract": "\\noindent The observational data on the large scale structure (LSS) of the Universe are used to establish the upper limit on the neutrino content marginalized over all other cosmological parameters within the class of adiabatic inflationary models. It is shown that the upper 2$\\sigma$ limit on the neutrino content can be expressed in the form $\\Omega_{\\nu}h^2/N_{\\nu}^{0.64}\\le0.042$ or, via the neutrino mass, $m_{\\nu}\\le4.0$eV. ", "introduction": " ", "conclusions": "" }, "0011/astro-ph0011325_arXiv.txt": { "abstract": "We present a new analysis of the currently available orbital elements for the known Kuiper belt objects. In the non-resonant, main Kuiper belt we find a statistically significant relationship between an object's absolute magnitude ($H$) and its inclination ($i$). Objects with $H<6.5$ (i$.$e$.$ radii $\\gapprox 170$km for a 4\\% albedo) have higher inclinations than those with $H>6.5$ (radii $\\lapprox 170\\,$km). We have shown that this relationship is not caused by any obvious observational bias. We argue that the main Kuiper belt consists of the superposition of two distinct distributions. One is dynamically hot with inclinations as large as $\\sim 35^\\circ$ and absolute magnitudes as bright as $4.5$; the other is dynamically cold with $i\\lapprox 5^\\circ$ and $H>6.5$. The dynamically cold population is most likely dynamically primordial. We speculate on the potential causes of this relationship. ", "introduction": "\\label{sec_intro} The discovery of the Kuiper belt in 1992 (Jewitt \\& Luu~1993) issued in a new era for the study of the outer solar system. The Kuiper belt is important not only because it is a rich, new region of the solar system to be explored, but because it contains important fossil clues about the formation of the outer solar system in particular, and about planet formation in general. Since its discovery, the Kuiper belt has supplied us with surprise after surprise. For example, before it was discovered, theorists believed that the Kuiper belt would consist of objects on low-inclination, nearly-circular orbits beyond the orbit of Neptune (Levison \\& Duncan~1993; Holman \\& Wisdom~1993). This belief seemed to be confirmed with the discovery of the first two Kuiper Belt Objects (hereafter KBOs), 1992~QB$_1$ and 1993~FW. However, the next four objects discovered revealed a real surprise. At the time of discovery their heliocentric distances were close enough to Neptune's orbit that their orbits should be unstable, unless protected by some dynamical mechanism. Indeed, many believed that they might have been Neptunian Trojans. However, these were the first discoveries of an unexpected population of objects on highly eccentric (up to 0.3) orbits in the 2:3 mean motion resonance with Neptune (co-orbiting with Pluto). Currently, objects in the trans-Neptunian region are divided into two main groups (see Malhotra et al$.$~2000 for a review). The {\\it Kuiper belt} consists of objects that are primarily on long-lived orbits, while the {\\it scattered disk} consists of objects that have suffered a close encounter with Neptune (Duncan \\& Levison~1997; Luu et al$.$~1997). The Kuiper belt itself is typically subdivided into two populations. Inside of roughly $42\\au$, objects tend to be locked into mean motion resonances with Neptune. Most known objects in this class are in Neptune's 2:3 mean motion resonance. However, a fraction also reside in the 3:5 and the 3:4 resonances. The orbits of all these objects are probably a result of resonance capture during the slow outward migration of Neptune during the late stages of planet formation (Malhotra~1995). Beyond $42\\au$, although several objects are believed to be in the 1:2 mean motion resonance (Marsden~2000a), most objects are not on resonant orbits. These non-resonant objects are members of what has come to be called the {\\it main Kuiper belt}. Models of planetary migration (e$.$g$.$ Malhotra~1995; Holman~1995; Hahn \\& Malhotra~1999) predict that unlike the KBOs in mean motion resonances, main KBOs should be on relatively low-inclination, nearly-circular orbits. However, recent observations have shown that this is not the case. Numerous objects in this region have very large inclinations\\footnote{Eccentricities are not a good measure of how excited the Kuiper belt is since most large eccentricity orbits are removed through close encounters with Neptune, truncating the eccentricity distribution. Inclinations do not suffer from this problem (Duncan, et al$.$~1995).}, certainly up to about $32^\\circ$, and most likely even higher (Marsden~2000a). Several papers have been published which attempt, among other things, to explain the high inclinations seen in the main Kuiper belt. The mechanisms invoked to date involve the scattering of KBOs by large objects temporarily evolving through the region. It takes a massive object to excite KBOs to high inclination; much more massive than the KBOs themselves\\footnote{A simple calculation based on an object's escape velocity shows that it must be larger than roughly twice the radius of Pluto to scatter a Kuiper belt object to an inclination of 30$^\\circ$.}. Petit et al$.$~(1999) suggest that the dynamically excited Kuiper belt is caused by the passage of Earth-mass objects through that region of the solar system. Thommes et al$.$~(1999) suggest that the large inclinations are due to the passage of Uranus and/or Neptune through the Kuiper belt while on eccentric orbits, after these planets were ejected from the region between Jupiter and Saturn. Ida et al$.$~(2000) suggest that the Kuiper belt was excited by a passing star. In this paper we present an analysis of the currently available orbital data of main belt KBOs which shows a new and surprising trend --- an unexpected and intriguing correlation between inclination and absolute magnitude. In particular intrinsically bright objects tend to be found on larger inclinations than do intrinsically faint objects. In \\S{\\ref{sec_data}} we present the data and discuss the statistical significance of this trend. In \\S{\\ref{sec_bias}} we investigate whether this trend is a result of observational selection effects. Our preliminary interpretation of this trend is presented in \\S{\\ref{sec_interp}}. We summarize our findings in \\S{\\ref{sec_concl}}. \\clearpage ", "conclusions": "\\label{sec_concl} We have shown that the inclination distribution of objects in the main Kuiper belt most likely varies as a function of absolute magnitude. In particular, objects intrinsically brighter than $H=6.5$ appear to have systematically higher inclinations than intrinsically fainter objects. There is only $\\sim 3\\%$ chance that these two distributions are the same. We have shown that this result is unlikely to be caused by biases in discovery or recovery observing procedures. Therefore, although it is possible that this conclusion is a result of small number statistics, we believe that it is real. Future discoveries and followups will clearly resolve this issue. The clear implication of our result is that a main belt object's inclination is dependent on its size. The differences between intrinsically bright objects and the intrinsically faint objects is best seen in Figure~\\ref{fig_iH}. Perhaps the most natural interpretation for the data in this figure is that we are seeing the superposition of two distinct populations. The first contains a dynamically hot population (inclinations up to $\\sim 35^\\circ$) consisting of both large and small objects (absolute magnitudes as small as 4.5 or radii up to $\\sim 330\\,$km for albedos of 4\\%). Indeed, even larger objects and/or objects with higher inclinations are likely to still be found. The other population is a dynamically cold one ($i\\lapprox 5^\\circ$) preferentially containing smaller objects ($H\\gapprox6.5$ or radii $\\lapprox 170\\,$km for albedos of 4\\%)." }, "0011/astro-ph0011113_arXiv.txt": { "abstract": "We develop a method for recovering the global density distribution of the ancient Galactic stellar halo prior to disk formation, based on the {\\it present} orbits of metal-poor stars observed in the solar neighborhood. The method relies on the adiabatic invariance of the action integrals of motion for the halo population during the slow accumulation of a disk component, subsequent to earlier halo formation. The method is then applied to a sample of local stars with [Fe/H]$\\le-1.5$, likely to be dominated by the halo component, taken from Beers et al.'s recently revised and supplemented catalog of metal-poor stars selected without kinematic bias. We find that even if the Galactic potential is made spherical by removing the disk component in an adiabatic manner, the halo density distribution in the inner halo region ($R \\le 15$ kpc) remains moderately flattened, with axial ratio of about 0.8 for stars in the abundance range [Fe/H]$\\le -1.8$ and about 0.7 for the more metal-rich interval $-1.8<$[Fe/H]$\\le-1.5$. The outer halo remains spherical for both abundance intervals. We also find that this initial flattening of the inner halo is caused by the anisotropic velocity dispersions of the halo stars. These results suggest that the two-component nature of the present-day stellar halo, characterized by a highly flattened inner halo and nearly spherical outer halo, is a consequence of {\\it both} an initially two-component density distribution of the halo (perhaps a signature of dissipative halo formation) {\\it and} of the adiabatic flattening of the inner part by later disk formation. Further implications of our results for the formation of the Galaxy are also discussed, in particular in the context of the hierarchical clustering scenario of galaxy formation. ", "introduction": "The currently favored cold dark matter (CDM) theory of galaxy formation postulates that the formation of a massive spiral galaxy like our own is a consequence of the hierarchical assembly of subgalactic dark halos, and the subsequent accretion of cooled baryonic gas in a virialized, galaxy-scale dark halo (e.g., Peacock 1999). Numerical studies based on this picture are able to, at least qualitatively, reproduce the characteristic features of a disk galaxy -- the massive dark halo, the stellar halo, and the stellar disk components (e.g., Steinmetz \\& M\\\"uller 1995; Bekki \\& Chiba 2000; Navarro \\& Steinmetz 2000), though difficulties are still encountered in the details. For example, the simulations conducted to date do not adequately account for the size of the disk component and the number of satellite galaxies (Navarro, Frenk, \\& White 1995; Moore et al. 1999; Klypin et al. 1999). The CDM hierarchical model may be regarded, in its essence, as a modern generalization of the classical Searle \\& Zinn (1978) hypothesis for the formation of the Galactic stellar halo. To explain a large inferred spread in the ages of globular clusters, and the lack of a spatial gradient in their metal abundances, Searle \\& Zinn argued that the halo component may have experienced prolonged, chaotic accretion of subgalactic fragments, as opposed to the rapid, monolithic collapse proposed by Eggen, Lynden-Bell, \\& Sandage (1962). Recent discoveries of halo substructures in Galactic phase space (Majewski, Munn, \\& Hawley 1994; 1996; Helmi et al. 1999; Chiba \\& Beers 2000, hereafter CB; Yanny et al. 2000) and of the Sagittarius dwarf galaxy, which is presently being disrupted by the Galactic tidal field (Ibata, Gilmore, \\& Irwin 1994; Ibata et al. 2000), may lend further support to this picture. Although halo formation via hierarchical assembly of subgalactic systems, such as dwarf galaxies, may continue to the present day (Bland-Hawthorn \\& Freeman 2000), a large fraction of the stellar halo, especially the inner part where the disk lies, should have been completed {\\it prior to disk formation}, since otherwise the disk component is made significantly thicker than is observed due to dynamical heating from infalling masses (Toth \\& Ostriker 1992). A clear age gap between the (thin) disk and the stellar halo supports that the latter consists of ancient populations (e.g. Liu \\& Chaboyer 2000). Also, studies of star-forming histories in the disk component indicate that the disk has been accumulated at an approximately constant rate over the last several billion years (Twarog 1980; Sommer-Larsen \\& Yoshii 1990); frequent mergings of dwarf galaxies over the Galaxy's lifetime will entirely modify the photometric and spectroscopic properties of the disk. Thus, one may well postulate that the inner part of the stellar halo, at say $R \\le 15$ kpc, retains a fossil imprint of how it was formed. The question then arises, what {\\it was} the structure of the halo component prior to disk formation? Since the bulk of halo stars are found in this inner part, where the disk gravity is dominant, the present-day structure of the halo can be greatly affected by later disk formation. It is thus necessary to consider the dynamical effect of the disk in inferring the structure of the ancient halo from the currently observed halo stars. Binney \\& May (1986, hereafter BM) examined this issue by assuming adiabatic invariance (for halo stars) of the action integrals of motion ${\\bf J}$, and for the distribution function $f({\\bf J})$ during slow disk formation. They set up test particles distributed in a spheroid, similarly to the stellar distributions of elliptical galaxies, and calculated the dynamical response of the particles to the slow increase of disk mass inside the spheroid. They showed that the Galactic halo, before the disk was formed, may have had a somewhat flattened density distribution (axial ratio $q \\sim 0.7$), in order to produce a current highly flattened halo ($q \\sim 0.3$), which they inferred from the radially anisotropic velocity ellipsoid of local metal-poor stars. We note here that recent kinematic data for larger samples of local metal-poor stars indicate a more moderately flattened halo ($q \\sim 0.7$) for inner radii ($R \\le 15$ kpc), whereas the outer halo is nearly spherical (Sommer-Larsen \\& Zhen 1990, hereafter SLZ; CB). This is also supported by examination of the spatial distributions of other halo tracers (e.g., Hartwick 1987; Preston et al. 1991; Kinman, Suntzeff, \\& Kraft 1994; Yanny et al. 2000). Also, the extent to which the formation of the disk component flattened the halo depends on the unknown initial velocity distribution of halo stars, whereas the initial conditions set up by BM apply only for one specific case. Thus, it is yet unexplored what the currently available data for metal-poor stars may tell us about the structure of the ancient halo before the disk was formed. In this paper we revisit this issue, based on a large sample of halo stars in the solar neighborhood, taken from a recently completed catalog of metal-poor stars selected without kinematic bias (Beers et al. 2000). It is noted that in a similar vein, Sommer-Larsen (1986), in his thesis work, arrived at a conclusion similar to BM's by investigating the distribution of individual orbital inclinations for his sample of 143 stars with [Fe/H]$\\le-1.2$. In contrast, we seek herein to develop a more general and direct method, based on the BM picture, to calculate the global density distribution of the halo prior to disk formation. The method is then applied to the more accurate and numerous data for metal-poor stars that is presently available. This paper is organized as follows. In \\S 2 we describe general properties of orbits in the mass model of the St\\\"ackel type that we adopt here, as well as the methodology for constructing the global density of a given local sample both before and after disk formation. The mass model for the Galactic potential that we adopt, and the local sample of metal-poor stars used in the current analysis, are also described. In \\S 3 we compute the orbital motions of the sample stars while conserving action integrals of motion. We then present the adiabatic change of the derived density distributions and kinematic properties of the sample when the disk component is slowly removed (essentially working backwards from the present to the past). Finally, in \\S 5, the results are summarized, and their implications for the formation and evolution of the Galaxy are discussed. ", "conclusions": "The global structure of the present-day stellar halo is characterized by an inner, highly flattened part, as revealed at $R<15$ kpc, and an outer, nearly spherical part (SLZ; CB). This two-component picture for the present-day stellar halo provides a reasonable explanation why faint-star-count studies have generally yielded an approximately spherical halo, whereas the local anisotropic velocities of the halo stars suggest a highly flattened system (Freeman 1987). The issue relevant here is what physical mechanism in the early stage of the Galaxy gives rise to the inner, highly flattened halo, where the bulk of halo stars are found, and where the effects of later satellite accretion may be diminished. One of the possible reasons for the two-component nature of the present-day halo is the slow formation of the disk within an initially spherical stellar halo (BM). In this paper, we have quantified the effect of later disk formation on the halo flattening, based on methods assuming adiabatic invariance of the motion of halo stars, and its application to a large sample of stars in the solar neighborhood. We have found that, even before disk formation, the inner part of the stellar halo exhibited a finite flattening, although it is more moderate than presently observed. The axial ratios of the density profiles within the almost spherical potential are $q \\simeq 0.80$ for [Fe/H]$\\le-1.8$ and $q \\simeq 0.70$ for $-1.8<$[Fe/H]$\\le-1.5$. Also, the initial velocity dispersions are characterized by an anisotropic velocity ellipsoid, as $(\\sigma_\\phi/\\sigma_\\lambda,\\sigma_\\nu/\\sigma_\\lambda)=(0.77,0.60)$ for [Fe/H]$\\le-1.8$ and (0.59,0.44) for $-1.8<$[Fe/H]$\\le-1.5$ at $R=R_\\odot$. Therefore, the inner part of the stellar halo was flattened by velocity anisotropy, and the more metal-rich population likely exhibited a more flattened density distribution. Through a comparison of the inclination angles of orbits for two abundance ranges, [Fe/H]$\\le-1.5$ and $-1.5<$[Fe/H]$\\le-1.2$, Sommer-Larsen (1986) also obtained a more flattened initial distribution for more metal-rich populations. We note that his latter subsample shows a rapid systematic rotation of 147 km s$^{-1}$, possibly contaminated by the stars belonging to the metal-weak thick disk (Freeman 1987). In contrast, our metal-rich halo subsample is selected from the more restrictive range $-1.8<$[Fe/H]$\\le-1.5$, where the effect of disk-like kinematics is minimal (Chiba \\& Yoshii 1998; CB) -- the flattening of this subsample is caused by velocity anisotropy, not systematic rotation. The results presented here may suggest that the ancient halo, at least in its inner part, may have undergone a somewhat ordered contraction. This contraction could have involved dissipation due to baryonic gas -- radiative cooling of this gas was most efficient in the innermost regions with high density. The resultant contraction of this gas, as a whole, may have proceeded mainly along the axis of rotation, because of the absence of the angular-momentum barrier in this direction. As the chemical enrichment proceeded along with the progress of the collapse, more metal-enriched stars would have ``seen'' a more flattened density distribution. On the other hand, the outer part of the halo may have been more susceptible to later infall of satellite galaxies, so that its density, kinematics, and mean age are different from those of the inner halo (Norris 1994; Carney et al. 1996; Sommer-Larsen et al. 1997). Alternatively, one might argue that the two-component nature of the present-day stellar halo is entirely a consequence of satellite accretion {\\it after} disk formation (Freeman 1987). According to Quinn \\& Goodman (1986) (see also Quinn, Hernquist, \\& Fullagar 1993), prograde satellite orbits that are initially inclined at less than about 60$^\\circ$ to the disk are dragged down quickly toward the plane by the effects of the dynamical friction of the disk. Walker, Mihos, \\& Hernquist (1996) further explore the effects of mergers of small satellites with large disk galaxies such as the Milky Way. The debris from these merging satellites, in combination with disrupted disk stars, would be expected to form a flattened system. Furthermore, if more massive satellites were more metal-rich (as they appear to be at present, see Mateo 1998), their orbits would fall farther toward the Galactic center so that their debris would form a more flattened, more metal-rich system (Freeman 1987). However, there exists no clear evidence for the predicted dynamical heating of the thin-disk component -- its very thin geometry (Toth \\& Ostriker 1992), and the nearly constant velocity dispersion of thin-disk stars over the last 10 Gyrs (Quillen \\& Garnett 2000), suggest that the thin disk has sustained little significant damage since its formation. In this regard, one might argue that the metal-weak thick disk is evidence for early dynamical heating of a pre-existing, metal-deficient, thin disk. However, it is then difficult to explain its absence in the abundance range considered in this paper (Chiba \\& Yoshii 1998; CB). It is worthwhile to remark that the hypothesis of the dissipative formation of the inner flattened halo, as well as the later accretion of satellites onto the outer halo, is a natural consequence of the CDM hierarchical clustering model (Bekki \\& Chiba 2000). This model postulates that a protogalactic system initially contains numerous subgalactic clumps, comprised of a mixture of gas and dark matter, and that the merging of these clumps led to a smaller number of more massive clumps. In the simulations of Bekki \\& Chiba (2000), these larger clumps move gradually toward the central region of the system, due to both dynamical friction and dissipative merging with smaller clumps. Finally, the last merging event occurs between the two most massive clumps, and the metal-poor stars which have been formed inside the clumps are disrupted and spread over the inner part of the halo. The aftermath is characterized by a flattened density distribution. Some fraction of the disrupted gas from the clumps may settle into the central region of the system, and produce a more enriched, more flattened density distribution. Some of the initially small density fluctuations in the outer region would have gained systematically higher angular momentum from their surroundings, and then slowly fallen into the system after most parts of the system were formed. This may correspond to the process of late satellite accretion, contributing primarily to the outer part of the halo. Thus, the reported initial state of the stellar halo can be explained, at least qualitatively, in the context of hierarchical clustering scenario. An alternative approach for elucidation of the dissipative nature of halo formation is to examine the results of recent high-resolution N-body simulations of structure formation based on the CDM theory (e.g., Ghigna et al. 1998; Moore et al. 1999; Klypin et al. 1999). Such simulations provide the orbital properties of dark matter particles inside virialized dark halos. If the stellar halo component in the Galaxy is formed similarly through dissipationless hierarchical assembly, the orbits of halo stars prior to disk formation, as derived in the current paper, may follow those of dark matter particles. For this purpose, we take the Ghigna et al. (1998) simulation of the formation of a cluster, as this is currently the only published one that presents the detailed orbital distribution of the simulated particles. They showed that the orbital distribution of the halo particles is close to isotropic -- circular orbits are rare and radial orbits are common. The average ratio of pericentric and apocentric distances, $r_{pr}/r_{ap}$, is equal to or less than 0.20, without showing a large variation as a function of the distance from the cluster: the median ratio is approximately 0.17. On the other hand, as is deduced from Figure 2, the orbits of the halo stars we have derived here before the disk was in place are more circular than the simulated dark halo particles, and the velocity field is anisotropic: the average value of $r_{pr}/r_{ap}$ is 0.29. Thus, we require some additional process, possibly dissipative interaction among protogalactic clumps to circularize their orbits, to explain the characteristic orbital distribution of the halo stars in the early Galaxy. More quantitative conclusions must await more elaborate modeling of the formation of the Galaxy over a large number of possible model parameters. Also, it is necessary to assemble and analyze the data of more remote stars, especially those presently found inside the solar radius, where our modeling of the halo is incomplete. In this regard, the next generation of astrometric satellites, such as {\\it FAME} and {\\it GAIA}, will provide highly precise parallaxes and proper motions for numerous stars, so that both three dimensional positions and velocities will be available over a large fraction of the halo. Also, with these astrometric satellites, we will be able to determine the exact mass distributions of the disk and dark halo components, and thus obtain definite information on the early Galaxy, using the technique outlined here. Furthermore, in addition to the Milky Way, direct identification of halo populations in external disk galaxies may prove promising as a way to clarify the global structures of stellar halos and their association with disks and bulges (e.g., Morrison 1999). Such studies should be eagerly pursued with 10m class telescopes." }, "0011/astro-ph0011439_arXiv.txt": { "abstract": "I consider the growth of inhomogeneities in a low-density baryonic, vacuum energy-dominated universe in the context of modified Newtonian dynamics (MOND). I first write down a two-field Langrangian-based theory of MOND (non-relativistic), which embodies several assumptions such as constancy of the MOND acceleration parameter, association of a MOND force with peculiar accelerations only, and the deceleration of the Hubble flow as a background field which influences the dynamics of a finite size region. In the context of this theory, the equation for the evolution of spherically symmetric over-densities is non-linear and implies very rapid growth even in a low-density background, particularly at the epoch when the putative cosmological constant begins to dominate the Hubble expansion. Small comoving scales enter the MOND regime earlier than larger scales and therefore evolve to large over-densities sooner. Taking the initial COBE-normalized power spectrum provided by CMBFAST (Seljak \\& Zeldarriaga 1996), I find that the final power-spectrum resembles that of the standard $\\Lambda$CDM universe and thus retains the empirical successes of that model. ", "introduction": "A primary motivation for cosmic non-baryonic dark matter with negligible pressure is the necessity of forming the presently observed structure in the Universe without violating the constraints on temperature fluctuations in the CMB. Basically, this is because structure in the dark matter component on galaxy to super-cluster scales can begin growing via gravitational instability considerably before hydrogen recombination (Peebles 1982, Vittorio \\& Silk 1984, Bond \\& Efstathiou 1984). This remains one of the powerful arguments against a low density baryonic universe. Any alternative cosmology not including CDM must invoke some mechanism other than conventional gravitational collapse in order to form structure. McGaugh (1999) suggests that the modified Newtonian dynamics (MOND), proposed by Milgrom (1983) as an alternative to dark matter on on galaxy and cluster scales, can provide the needed mechanism and, further, that the consistency of the observed angular structure of the temperature fluctuations in the CMB (Lange et al. 2000, Hanany et al. 2000), with a pure baryonic universe (McGaugh 2000), may be viewed as support for MOND. This speculation is based upon the general expectation that MOND, in providing stronger effective gravity in the limit of low accelerations, would assist in structure formation. MOND is an {\\it ad hoc} modification of Newton's law of inertia or gravity at low acceleration. The original idea is contained in the statement that when the acceleration falls below $a_o$, a new physical constant with units of acceleration, then the effective gravitational acceleration approaches $\\sqrt{g_na_o}$ where $g_n$ is the usual Newtonian gravitational acceleration. Although this simple formula works remarkably well in describing galaxy rotation curves consistently with the observed distribution of detectable matter (Sanders 1996, McGaugh and de Blok 1998), it clearly lacks the generality to treat the problem of cosmological density fluctuations. A more consistent physical description of modified dynamics is provided by the non-relativistic Langrangian-based theory of Bekenstein and Milgrom (1984, hereafter BM). An obvious procedure, when treating the growth of density fluctuations, would be to take the modified Poisson equation of BM and consider small fluctuations about a zeroth order solution as in Newtonian cosmology. The problem is that, when applied to a finite sphere as is usual in Newtonian cosmology, the zeroth order solution is not that of a linear Hubble flow-- the absolute distance cannot be factored out and it is not possible to describe cosmology in terms of a universal scale factor. The cosmology is basically that described by Felten (1984) and Sanders (1998), in which MOND alters the usual Friedmann solutions; as soon as the cosmic deceleration over some physical scale falls below $a_o$, then that entire region begins to deviate from uniform Hubble flow. This leads to the eventual re-collapse of any finite size region regardless of its original density or expansion velocity. In this picture density fluctuations play no role. Apart from problems in principle (what determines the point or points about which MOND collapse proceeds?), this cosmology leads to clear contradictions with observations-- recollapse in the present Universe occurs out to scales of 30 Mpc. One might expect that in a proper theory, the basic Hubble flow remains intact, and structure develops from the field of small density fluctuations as in standard gravitational collapse. In order to construct a reasonable MOND cosmology which has this attribute, one must supplement the BM theory with several assumptions which may reasonably follow from a more general theory. The first of these assumptions-- also an aspect of the earlier MOND cosmology-- is that the MOND acceleration parameter, $a_o$, which is comparable to the acceleration in the outer regions of galaxies ($\\approx 10^{-8}$ cm/s$^{-2}$), does not vary with cosmic time. Numerically, $a_o \\approx cH_o/6$ which suggests that modified dynamics may reflect the influence of cosmology on local particle dynamics. If $a_o$ varies as the Hubble parameter, then the argumentation presented here would be incorrect. However, it is also possible that $a_o$ is related to the cosmological constant (Milgrom 1999) and is independent of cosmic time. If this is true then MOND plays no role in the evolution of the early radiation-dominated Universe since cosmic deceleration greatly exceeds $a_o$ on relevant scales (e.g. the Jeans length). In the later matter-dominated, pressureless evolution, the cosmic deceleration on co-moving scales corresponding to galaxies or clusters falls below $a_o$ and one might expect modified dynamics to affect the formation of such structure. The second assumption directly concerns the problem of the zeroth order Hubble flow; we wish to construct a theory in which MOND plays no role in the absence of fluctuations, and the background cosmology is essentially unaltered. In other words, MOND should apply only to peculiar accelerations-- the accelerations developing from inhomogeneities-- and not to the overall Hubble flow; i.e., no MOND in a homogeneous Universe. This assumption can find some justification in the context of a stratified scalar-tensor theory in which MOND phenomenology results from a scalar force that becomes dominant in the limit of low scalar field gradients (Sanders 1998). The third assumption concerns the influence of the Hubble flow on the internal dynamics of an otherwise isolated spherical region. In modified dynamics, and any covariant extention thereof, it must be the case that the internal dynamics of a sub-system is influenced by the presence of an external field-- the ``external field effect\" (Milgrom 1983). This is essentially an observational requirement on MOND imposed by the absence of discrepancies in Galactic star clusters. In other words, the underlying theory should not respect the equivalence principle in its strong version. With respect to cosmology, it is not clear how the external field effect would come into play, but I assume here that, for an over- or under-dense spherical region, the deceleration or acceleration of the Hubble flow is the one and only external field which influences the development of the inhomogeneity. Because the de-acceleration of the Hubble flow increases linearly with scale, fluctuations on small comoving scales are affected by MOND earlier than those on larger scale. One might expect this to lead a hierarchical scheme of structure formation, with smaller objects forming first. ", "conclusions": "Here I show that MOND provides the possibility of overcoming the problem of structure growth in a low-density baryonic Universe. In the context of the simple two-field non-relativistic theory of modified dynamics presented here, we see that when the background deceleration of the Hubble flow over a given scale falls below the critical MOND acceleration, $a_o$, then the growth of structure on that scale is greatly enhanced relative to the Newtonian expectation. The growth of over-densities on smaller scale is even more enhanced due to the fact that smaller regions enter the MOND regime earlier; the early growth of small-scale fluctuations can compensate for the effect of Silk damping on these scales. Thus the resulting power spectrum, apart from the oscillations, closely resembles that of the favored $\\Lambda$CDM cosmology. On the scale of galaxies (1.5 Mpc), even though the typical initial over-density is on the order of $2\\times 10^{-10}$, the fluctuation grows to the non-linear regime by a redshift of 2.5. Thus MOND would appear not only to explain the observed large scale structure, but also provide a mechanism for early galaxy formation. This is all achieved with the value of $a_o$ determined from galaxy rotation curves. The minimalist MOND theory has not been fine-tuned in any sense to match the observed power spectrum; the single adjustable parameter $\\beta$ lies within the range which is consistent with the observed form of galaxy rotation curves. Of course, these conclusions depend upon the approximate validity of the assumed theory described in section 2. This theory, and assumptions embodied therein, guarantee that the early, radiation-dominated evolution of the universe is identical to that of the standard model, that the basic Hubble flow is unaffected by MOND, and that fluctuations on the scale of galaxies to super-clusters enter the MOND regime, determined by the background Hubble flow, sufficiently early (but not too early) to assure growth to the present amplitude. Because of the necessity of such an {\\it ad hoc} theory in the absence of a more fundamental covariant theory, it is perhaps premature to compare in detail the predicted power spectrum or peculiar velocities with observations. In particular, the oscillations (Fig.\\ 2) may not actually be evident in the evolved power spectrum due to non-linear aspects of the theory which are ignored here-- specifically, not only individual Fourier components but also over-dense spherical regions cannot be considered in isolation (larger scale peculiar accelerations contribute to the background field). The detailed results shown in Figs.\\ 1 and 2 should be taken as a demonstration that MOND, in a low-density baryonic universe, can provide a vigorous growth of fluctuations-- growth which is sufficiently rapid to lead to the large scale structure observed at the current epoch. Finally, I re-emphasize that the presence of a dynamically significant cosmological constant plays a necessary role in the rapid growth of structure with this version of modified dynamics. The MOND growth of inhomogeneities accelerates at the epoch when $g_b\\approx 0$ due to the dominance of the non-linear term in eq.\\ 28 (only possible with a cosmological constant comparable to ${H_o}^2$). This adds a new aspect to an anthropic argument originally given by Milgrom (1989): we are observing the Universe at an epoch when $\\Omega_\\Lambda$ has only recently emerged as the dominant term in the Friedmann equation because it is only then that structure formation proceeds rapidly. If the evidence in support of a baryonic-$\\Lambda$ universe continues with further observations of the CMB angular power spectrum, then some unconventional mechanism for the formation of structure must be invoked. Here it is evident that modified dynamics, with a well-documented success in explaining the kinematic observations of galaxies and clusters without dark matter, may also successfully address the problem of structure formation in a low-density baryonic universe. I am grateful to Art Wolfe, Eric Gawiser and Kim Griest for teaching me all about anisotropies in the CMB and the standard treatment of the evolution of density fluctuations. I thank Stacy McGaugh for stimulating this work, Arthur Kosowsky and Jacob Bekenstein for helpful criticisms, and the referee, Jim Peebles, whose numerous critical remarks led to a considerable improvement in the content and presentation of this paper. Finally I am most grateful to Moti Milgrom for his typically penetrating comments on possible MOND cosmologies. \\newpage" }, "0011/astro-ph0011462_arXiv.txt": { "abstract": "In this paper, the cosmic microwave background (CMB) anisotropy in a multiply-connected compact flat 3-torus model with the cosmological constant is investigated. Using the COBE-DMR 4-year data, a full Bayesian analysis revealed that the constraint on the topology of the flat 3-torus model with low-matter-density is less stringent. As in compact hyperbolic models, the large-angle temperature fluctuations can be produced as the gravitational potential decays at the $\\Lambda$-dominant epoch well after the last scattering. The maximum allowed number $N$ of images of the cell (fundamental domain) within the observable region at present is approximately 49 for $\\Omega_m=0.1$ and $\\Omega_\\Lambda=0.9$ whereas $N\\sim8$ for $\\Omega_m=1.0$ and $\\Omega_\\Lambda=0$. ", "introduction": "For a long time, cosmologists have assumed the simply connectivity of the spatial hypersurface of the universe. If it is the case, the topology of closed 3-spaces is limited to that of a 3-sphere if Poincar\\'{e}'s conjecture is correct. However, if we assume that the spatial hypersurface is multiply connected, then the geometry of spatially finite models can be flat or hyperbolic as well. The metric describes only the local geometry. We should be able to observe the imprint of the ``finiteness'' of the spatial geometry if it is multiply connected on scales of the order of the particle horizon or less, in other words, if we live in a ``small universe''. \\\\ \\indent For flat multiply connected models without the cosmological constant, various constraints using the COBE DMR data have been obtained \\cite{Sokolov,Starobinsky,Stevens,Oliveira1,Oliveira2,Levin1,Roukema}. Assuming that the initial power spectrum is scale-invariant ($n\\!=\\!1$) the suppression of the temperature fluctuations on scales beyond the typical size of the cell $L$\\footnote{$L$ can be defined as twice the diameter which is defined as the maximum of the minimum geodesic distance between two points over the space.} leads to a decrease in the large-angle power. For 3-torus models without the cosmological constant in which the cell (fundamental domain) is a cube, the constraint is $L>0.8 R_\\ast$ where $R_\\ast$ is the comoving radius of the last scattering surface(horizon radius) \\cite{Sokolov,Starobinsky,Stevens,Oliveira1}. It should be emphasised that the constraint itself \\ti{does not imply that the ``small universe'' is ruled out} since the possible maximum expected number of the copies of the cell within the observable region at present is approximately 8. For models in which one side of the cell is longer than the others, the constraint for the smallest topological identification scale(=the diameter of the smallest ball which can wrap around the space) can be less stringent\\cite{Roukema}. \\\\ \\indent On the other hand, recent observations of distant supernova Ia \\cite{Perlmutter,Riess} imply the existence of ``missing energy `` which possesses a negative pressure and equation-of-state ($w\\equiv p/\\rho$) in the form of either a cosmological constant, vacuum energy density or a slowly-varying spatially inhomogeneous component ``quintessence''. Together with the first observation of the height and the position of the first Doppler peak by {{\\sc Boomerang}}\\cite{Boomerang} and {{\\sc MAXIMA}} \\cite{MAXIMA} and the COBE-DMR data, the constraint for the cosmological constant is $0.69<\\Omega_\\Lambda<0.82$ and for the matter(cold dark matter, baryon plus radiation) $0.28<\\Omega_m<0.42$ assuming adiabatic initial perturbations\\cite{Balbi}. \\\\ \\indent For low-matter-density models with flat geometry a bulk of large-angle cosmic microwave background(CMB) fluctuations can be produced as the gravitational potential decays at the $\\Lambda$-dominant epoch $1+z\\sim (\\Omega_\\Lambda/\\Omega_0)^{1/3}$. Recent works \\cite{Aurich,Inoue2,CS,AS} have shown that the angular power spectrum $C_{l}$ is completely consistent with the COBE-DMR data for some compact hyperbolic models which are incompatible with the previous analyses \\cite{Bond1}. Because the angular sizes of fluctuations produced at the late epoch are large compared to those on the last scattering for flat or hyperbolic geometry, we expect that the constraints for compact flat models with low-matter-density can be also significantly loosened. \\\\ \\indent In this paper, the CMB anisotropy in a multiply-connected compact flat 3-torus model with or without the cosmological constant is investigated. In sec II we briefly describe the time evolution of the scalar perturbation in the locally flat Friedmann-Robertson-Walker models which will be used for computing the CMB anisotropy. In sec III we compare the angular power spectrum in a flat 3-torus model with the cosmological constant to the one in the ``standard'' 3-torus model with $\\Omega_{tot}\\!=\\!1$. In sec IV a full Bayesian analysis using the COBE-DMR data has been carried out for giving the constraint on the minimum size of the cell. \\\\ \\indent ", "conclusions": "In this paper the CMB anisotropy in a flat 3-torus model with or without the cosmological constant has been investigated. Using the COBE-DMR data, we have done a full Bayesian analysis incorporating the effect of anisotropic correlation. It has turned out that the constraint on the low-matter-density model is less stringent compared to the ``standard'' model with $\\Omega_m\\!=\\!1.0$. The reason is that the large-angle fluctuations (on COBE scales) are produced at late time well after the last scattering. The physical size of these fluctuations is of the order of the size of the cell or less. Hence the effect of the non-trivial topology becomes insignificant. We expect that the result does not significantly change for other compact flat models with different topology\\cite{Levin1} or with different shape of the cell since the physical effect does not change. \\\\ \\indent We have seen that the analysis using the angular power spectrum is not enough since the background geometry is globally anisotropic. Even if the power spectrum well agrees with the data, it does not guarantee the validity of the model since the power spectrum has information of only isotropic components in the 2-point correlations. If the background geometry is globally anisotropic then the fluctuations form an anisotropic Gaussian field for a given orientation. On the other hand, the fluctuations become non-Gaussian if one marginalises the distribution function over the orientation\\cite{Magueijo}. For locally Friedmann-Robertson-Walker models which are spatially compact, we expect to observe zero skewness but non-zero kurtosis provided that the initial fluctuation is Gaussian. \\\\ \\indent However, for low-matter density models, anisotropy in the statistically averaged correlation on large angular scale cannot be so large since the physical size of observed temperature fluctuations that has been prodeced well after the last scatteing is much smaller than the actual size of the space(i.e. topological identification scale). \\\\ \\indent In the case of flat topology, the discrete eigenmodes have ``regular'' features which lead to significant correlations in $a_{lm}$'s. Even if marginalised over the orientation, the correlations do not completely disappear. This contrasts with the case of hyperbolic topology in which the discrete eigenmodes are ``chaotic'' and the correlations in $a_{lm}$'s disappear if one takes an average over the position of the observer. \\\\ \\indent In order that we would be able to observe the periodic structure in the fluctuations for the 3-torus models, we need to have $L/(2 R_\\ast)<1$ where $R_\\ast$ is the comoving radius of the last scattering surface. Because the obtained constraints give $L/(2 R_\\ast)>0.40,0.22$ for $(\\Omega_m,\\Omega_\\Lambda)=(1.0,0) ,(0.1,0.9)$, respectively, we still have a great chance of the first detection of the non-trivial topology of the universe by the future satellite missions such as MAP and \\ti{Planck} which can survey the CMB in the full sky with high resolution by measuring some specific signatures (e.g. the circles test \\cite{Cornish2} or the non-Gaussianity \\cite{Inoue3})." }, "0011/astro-ph0011148_arXiv.txt": { "abstract": "We study static neutron stars with poloidal magnetic fields and a simple class of electric current distributions consistent with the requirement of stationarity. For this class of electric current distributions, we find that magnetic fields are too large for static configurations to exist when the magnetic force pushes a sufficient amount of mass off-center that the gravitational force points outward near the origin in the equatorial plane. (In our coordinates an outward gravitational force corresponds to $\\partial\\ln g_{tt}/\\partial r>0$, where $t$ and $r$ are respectively time and radial coordinates and $g_{tt}$ is coefficient of $dt^2$ in the line element.) For the equations of state (EOSs) employed in previous work, we obtain configurations of higher mass than had been reported; we also present results with more recent EOSs. For all EOSs studied, we find that the maximum mass among these static configurations with magnetic fields is noticeably larger than the maximum mass attainable by uniform rotation, and that for fixed values of baryon number the maximum mass configurations are all characterized by an off-center density maximum. ", "introduction": "\\label{sec:intro} Over the years, the typical magnitudes of the surface magnetic fields of pulsars---as inferred from measured spindown rates and simple magnetic dipole models---have been $\\sim 10^{12}-10^{13}$ G \\citep{tayl93}. Assuming magnetic flux conservation, fields of $\\sim 10^{12}$ G would naturally arise from typical main sequence star surface magnetic fields of $\\sim 100$ G during a decrease in radius by a factor of $\\sim 10^5$ \\citep{shap83}. At the extreme end of fields attainable by flux conservation, the largerst observed white dwarf magnetic field of $5\\times 10^8$ G leads to a neutron star field of $\\sim 1\\times 10^{14}$ G \\citep{carr96}, while the largest observed main sequence stellar magnetic field of $3.4 \\times 10^4$~G \\citep{bohm89} also suggests a possible field of a few times $10^{14}$~G. Several independent circumstantial arguments link the class of objects known as ``soft $\\gamma$-ray repeaters'' (SGRs), and perhaps the so-called ``anomalous X-ray pulsars'' (AXPs), with neutron stars having magnetic fields $\\gtrsim 10^{14}$ G---the so-called ``magnetars'' \\citep{dunc92,usov92,pacz92,thom95,thom96,vasi97a}. (Table \\ref{tbl-0} displays some observed properties of these objects.) In addition to the circumstantial arguments, more direct evidence for magnetic fields of $2-8\\times 10^{14}$ G is available for two of the five known SGRs, in the form of measured periods and spindown rates of associated X-ray pulsars \\citep{kouv98,kouv99}.\\footnote{For ongoing discussion of this interpretation of X-ray timing data see e.g. \\cite{mars99a,wood99a,hard99,mars99b,chat00}.} Furthermore, the observed X-ray luminosities of the AXPs may require a field strength $B\\gtrsim 10^{16}$ G \\citep{chat00,heyl98b}. The population statistics of SGRs suggest that magnetars may constitute a significant fraction ($\\gtrsim 10\\%$) of the neutron star population \\citep{kouv94,kouv98}. As mentioned above, there are isolated examples of progenitor stars which could yield fields of $\\sim 10^{14}$ G by flux conservation, but these isolated examples do not seem sufficient to account for a significant fraction of the neutron star population. Thus, it seems likely that some mechanism {\\em generates} magnetic fields in nascent neutron stars. For example, \\cite{dunc92} suggest that the smoothing out of differential rotation and convection could generate fields as large as $3\\times 10^{17} (P_i/1\\ \\rm{ms})^{-1}$ G, where $P_i$ is the initial rotation period of the neutron star. These considerations motivate study of the effects of ultra-strong magnetic fields on neutron star properties. In this we have been inspired by the pioneering work of Bocquet, Bonazzola, Gourgoulhon \\& Novak (1995), who performed relativistic calculations of axisymmetric neutron star structure in which the standard stress-energy tensors of a perfect fluid and the electromagnetic field were employed, and were comparable in magnitude. The maximum fields they found were of order $10^{18}$ G, with increases of 13 to 29\\% in the maximum mass of nonrotating stars for various equations of state. An additional motivation is provided by the recent findings that magnetic fields of strengths larger than $10^{16}$ G affect the EOS of dense matter directly through drastic changes in the composition of matter \\citep{cha96,cha97,yz99,bpl00}. The EOS is altered by both the Landau quantization of the charged particles (such as protons, electrons, etc.) and the interactions of the magnetic moments, including the anomalous magentic moments of the neutral particles (such as the neutron, strangeness-bearing $\\Lambda$-hyperon etc.) with the magnetic field. In this work, we consider only the effects of the magnetic field on the structure, through its influence on the metric, in order to facilitate a comparison with the earlier work of \\cite{bocq95}. The additional changes caused by the direct effects of the magnetic field on the EOS will be reported in a future work \\citep{card00}. While \\cite{bocq95} also presented solutions for rotating neutron stars endowed with large magnetic fields, in this work we present only static solutions. In terms of the potential observability of the effects of large magnetic fields, the most relevant situation appears to be for non-rotating stars. Neutron stars with the highest inferred magnetic fields---the so-called ``magnetars''---are all observed to be rotating very slowly, with periods on the order of seconds. The effects of such slow rotation should have a negligible impact on the neutron star structure. In this paper we extend the work of \\cite{bocq95}. The theoretical formalism is outlined in \\S\\ref{sec:form}, which serves to put the problem in context. In \\S\\ref{sec:himass} we shed light on an issue left somewhat unclear by \\cite{bocq95}: What physically determines the maximum mass and magnetic field for a given maximum density or given baryon mass? In order to explore these questions we have chosen to solve the structural equations using a Green's function technique rather than the spectral technique employed by \\cite{bocq95}, and we also searched for the maximum mass, for a given magnetic field distribution, in a different way. An appendix describes our numerical methods and tests of our code. In \\S\\ref{sec:constmb} we present an illuminating view of constant baryon mass and constant magnetic moment sequences, and present higher mass configurations than those found by \\cite{bocq95} for the equations of state (EOSs) they employed. In addition, we present the results of analogous calculations using three more recent EOSs. Summary and outlook are contained in \\S\\ref{sec:sumout}. ", "conclusions": "\\label{sec:sumout} In summary, we present a method of computing the structure of axisymmteric relativistic stars that combines elements of previous approaches, and report tests of our code. A quantitative method of determining the outer envelope (in the mass vs. radius plane) of configurations attainable with poloidal magnetic fields governed by a constant ``current function'' [see equation (\\ref{current})] has been found: magnetic fields are too large for static configurations to exist when the magnetic force pushes a sufficient amount of mass off-center that the gravitational force points outward near the origin in the equatorial plane. (In our coordinates an outward gravitational force corresponds to $-\\partial \\nu/\\partial r>0$.) We obtain larger masses of neutron stars in ultra-strong magnetic fields than have been reported previously for various equations of state, and performed computations with three representative modern EOSs. Sequences of constant baryon mass and constant magnetic moment are displayed. For all EOSs studied, the maximum attainable mass of static stars with a magnetic field determined by a constant current function is noticeably larger than that attainable with uniform rotation and no magnetic field. The results presented here are only an initial step in exploring possible configurations of neutron stars with strong magnetic fields. As we mention below, configurations with azimuthal field components will be of physical interest, which implies that three-dimensional geometries should also be considered. But even with attention restricted to poloidal fields, we have only scratched the surface of possible configurations, as we have only considered a single current function. \\cite{bocq95} make brief mention of computations with a few other current functions. We have performed a handful of exploratory computations using a polytropic EOS and other current functions and have found some toroidal solutions. These toroidal solutions were not attainable with the computational approach of \\cite{bocq95}, since their method involved the specification of a finite density at the center of the star. In the case of toroidal configurations, the simple condition determining the boundary of existence of static configurations will have to be generalized, since there is no matter at the center. These explorations will be reported in detail elsewhere. Our work here has focused on the effects magnetic fields have on general relativistic structure, ignoring the effects of intense magnetic fields on the EOSs. Recently, the direct effects of magnetic fields on the EOS have also been investigated \\citep{cha96,cha97,yz99,bpl00}. Substantial effects on the EOS above nuclear saturation densities are generally produced by fields in excess of $10^{18}$ G, which is of the order of the maximum central field strengths found in this paper. The generic effects on the EOS include softening due to Landau quantization, which is, however, overwhelmed by stiffening due to the incorporation of the magnetic moments of the various particles in neutron star matter \\citep{bpl00}. (Note that the important $B^2/8\\pi$ term is already included in our study.) Work is in progress (Cardall et al. 2000, in preparation) to provide fully self-consistent calculations of neutron star structure, in which the direct effects of magnetic fields on the EOS will be included in addition to the structural effects considered in this work. An issue which we defer to future work is the question of stability. For systems governed by a finite number of parameters, a generalization \\citep{sork82} of the familiar one-dimensional turning point method can be employed; see \\cite{frie88} for an application to uniformly rotating relativistic stars. However, as with differentially rotating stars, this generalized turning point method is not really applicable in the present case. This is because the need to specify a current function (or rotation law in the case of differential rotation) means that defining a particular configuration requires the specification of an {\\em infinite} number of parameters. Another issue that needs further explication before the physical relevance of the results presented here can be fully assessed is the generation of magnetic fields. We have mentioned the mechanism of \\cite{dunc92}, the generation of magnetic fields during the smoothing of differential rotation. However, the azimuthal dragging of field lines by differential rotation leads to nonvanishing azimuthal field components, in constrast to the poloidal fields studied here. It is not clear whether fields with azimuthal components would evolve into poloidal configurations, or whether there are mechanisms to directly generate strong poloidal fields. It would be of interest to explore the possibility of finding stationary solutions with toroidal magnetic fields, and in three dimensions. While this would involve more nonzero metric components, perhaps methods similar to those employed in this paper could be employed; see \\cite{bona98}. We wish to thank E. Gourgoulhon for helpful communications concerning the calculations of \\cite{bocq95}. We are grateful to Dany Page and Ralph Wijers for their help in the preparation of Table 1. Research support from DOE grants FG02-87ER40317 (for CYC and JML) and FG02-88ER-40388 (for MP) are gratefully acknowledged. \\appendix" }, "0011/gr-qc0011049_arXiv.txt": { "abstract": "{ We discuss scenarios in which the galactic dark matter in spiral galaxies is described by a long range coherent field which settles in a stationary configuration that might account for the features of the galactic rotation curves. The simplest possibility is to consider scalar fields, so we discuss in particular, two mechanisms that would account for the settlement of the scalar field in a non-trivial configuration in the absence of a direct coupling of the field with ordinary matter: topological defects, and spontaneous scalarization.} \\vskip 1.5cm \\noindent PACS number(s): 11.27.+d, 04.40.-b, 98.62.Gq \\newpage \\newpage ", "introduction": "It has been known for a long time that the motion of the stars and gases around the center of most galaxies can not be explained in terms of the luminous matter content of the galaxies, at least not within the context of Newtonian gravity (see \\cite{rubin} for a review). The standard view is that there is in almost every galaxy a large component of non-luminous matter (the gravitational dark matter) that forms an halo around the galaxy and that provides the additional gravitational attraction needed to explain the ``rotation curves'' in terms of standard gravitational theory. There are several proposals for this dark component, ranging from new exotic particles such as those predicted by supersymmetry \\cite{turner,primack,susy}, to other less exotic candidates such as massive neutrinos, all collectively known as WIMPs (weakly interacting massive particles) (see \\cite{turner,kolb} for a review), to the relative mundane idea of dark but ordinary bodies such as Jupiter-like objects collectively known as MACHOs (Massive Compact Halo Objects)\\cite{carr}. Searches for these types of objects have been made \\cite{alcock}, and although they report some findings, there doesn't seem to be enough of these objects to account for the galactic dynamics. Moreover there are severe bounds on the amount of baryonic matter in the universe arising from big bang nucleosynthesis and for some values of the Hubble constant those bounds also imply that some of the galactic dark matter ought to be exotic\\cite{peebles1,weinberg,copi}. Independently of this and despite their popularity, these type of models suffer from various problems and require surprising coincidences (see for example \\cite{sellwood}).\\\\ \\\\ Another type of proposal, which is in some sense more radical, is based on the idea that the gravitational theory would have to be modified when dealing with the scales associated with the motion of stars in galaxies \\cite{finzi,sanders}, in particular the idea is embodied by the proposal of Milgrom \\cite{milgrom}, that the laws of motion are modified when the accelerations involved are extremely small. Unfortunately this scenario has not, as yet, been converted into a fully relativistically invariant theory. Another type of model that has been exposed is to replace general relativity by a higher order in curvature theory, which in some particular cases appears to be obtaining encouraging results\\cite{Manheim}. The problem with this approach is that these types of theories have in general problems of principle like for example the lack of a well posed initial value formulation. Nevertheless such relatively radical proposals are still attractive, due in part to an intrinsic problems of the more conservative approaches in explaining the generality and universality of the phenomena, namely the fact that the amount of luminous matter seems to be such a good indicator of the amount of the dark matter component \\cite{persic} and the fact that the dark component happens always to distribute itself in such a way that the resulting rotation curves (hereafter referred to as RC) are almost flat away from the galactic centers \\cite{rubin1}. \\\\ \\\\ Thus, in contrast with the former scenarios which would need to assume not only the existence of the dark matter but also give some evolutionary scenarios that result in the aforementioned universality in its distribution, the modified gravity scenario would naturally account for such correlations without the need for additional assumptions. On the other hand the former scenarios do not present any problem in lending themselves to an acceptable theoretical formulation, compatible with present theories of particle physics and general relativity. \\\\ \\\\ The object of this article is to discuss a third type of scenario which has some of the advantages of each scheme. The idea is to take dark matter to be described not by a bunch of particles whose distribution needs to be explained but by a coherent field which would settle in a universal stationary configuration that would account for the generic features of the RC. The simplest possibility is provided by scalar fields, which would of course have to be very long ranged (i.e. masses smaller than ${1 \\over R_G}$ where $R_G$ is the radius of the largest galaxy with flat RC). The basic problem is that there are very severe experimental bounds for the direct coupling of such a field with ordinary matter \\cite{panov}, and in the absence of such coupling the field will in general settle globally in the minimum of the potential leading to a homogeneous configuration that will not produce the desired effects. On the other hand, one could hope that, given the likelihood of existence of large black holes at the center of most galaxies, they would account for the nontrivial configuration of the scalar fields. Unfortunately these kind of situation is largely forbidden by the ``Black Hole No hair theorems'' for scalar fields \\cite{MH,DS,JDB}. These limitations severely reduce the types of models one can consider, in particular, there are, known to these authors, only three mechanisms that would account for the settlement of the scalar field in a non-trivial configuration in the absence of a direct coupling of the field with ordinary matter (or some other exotic matter which we will not consider because of the incremental number of hypothesis it involves): a) boson-star like clumps, b) spontaneous scalarization, and c) topological defects. Other models that lack these features have been considered, for example in \\cite{tonatiuh}. However, such models face two problems: first, they give rise to spherical configurations where the scalar field in consideration is singular ``at the center'', and second, the resulting scalar field potential needed to account for the flat RC depends explicitly on the value of the ``tangential velocity'' of stars at the flat region. That is to say, such a potential have to be adjusted differently for different galaxies. Needless to say that both problems clearly make those schemes unsuitable as models for the problem at hand. \\bigskip Concerning the case ``a)'' mentioned above, it has been analyzed in \\cite{peebles}. Their analysis focuses on cosmological and evolutionary considerations as well as the issues related to the conditions under which the assumption of long range coherency of the scalar field is justified, rather than the universal features of the galactic rotation curves. We will deal here with the other two cases ``b)'' and ``c)''. \\\\ \\\\ The scenario ``b)'', namely the spontaneous scalarization (see Sec.V)\\cite{Damour}, is in some sense simpler because it involves a single scalar field in contrast of the various fields needed in the simplest versions of topological defects (e.g., global monopoles). Here the mechanism that allows for the nontrivial stationary configuration of the scalar field is connected to a non-minimal coupling of the scalar field to the curvature. This results in the effective gravitational coupling becoming dependent on the scalar field. The point is that such a coupling allows for the reduction of the total energy of the configuration (in comparison with the corresponding configuration with the same baryon number and no scalar field) for which the scalar field deviates from the trivial configuration by taking values that reduce the gravitational coupling in the regions of high matter density \\cite{SSN}. Thus the model must incorporate from the onset the non-minimal coupling that seems to be needed to account at the same time for the correlations in the dark-luminous matter components (see \\cite{NSS} and the discussion of the third scenario below). The disadvantage of this model, which is in fact shared by the first model (i.e., boson stars) \\cite{Igor}, is precisely the lack of resilience against black holes whose existence in most galaxies, if confirmed, would seem to preclude, through the no-hair theorems \\cite{MH,DS,JDB}, the models based on this mechanism. \\\\ \\\\ The scenario ``c)'' is exemplified by the model of global monopoles \\cite{vilenkin1} which have the notable feature of naturally leading to a $1/r^2$ energy density behavior which would naively account for the flat rotation curves and which upon taking the symmetry braking scale to be the GUT scale would result in the correct order of magnitude for the galactic dark matter. Unfortunately, upon further examination of the simplest model severe problems arise, in particular the monopole configuration turns out to be repulsive \\cite{harari}, and moreover the configuration would be too universal in the sense that it would be independent of the size of the galaxy thus defeating the hope for the correlation of dark to luminous matter over a range of galactic sizes. There is nevertheless hope to overcome these problems by the consideration of slightly more complicated models \\cite{NSS}. In that work the simple monopole model was supplemented by the introduction of a non-minimal coupling between the scalar fields and curvature (see \\cite{NSS} and Sec.VI). This resulted in the restoration of gravitational attraction leading to regions of relatively flat rotation curves and to the possibilities of the dark-luminous matter correlations arising from the fact that in these models the scalar potential $V(\\Phi^a \\Phi_a)$ (where $\\Phi^a$ stands for a triplet of scalar fields that characterize the global monopole) is replaced with the effective scalar potential $V(\\Phi^a\\Phi_a)+ F(\\Phi^a\\Phi_a,R)$ (here $R$ stands for the scalar curvature of the space-time metric) whose minima depend on the amount of matter present trough the effect of the latter on $R$. The global monopole model has the additional advantage of resilience against the formation of black holes in the galactic centers, since their topological charge makes them immune to the devastating limitations imposed by the no hair theorems. \\\\ \\\\ Despite the promising features of the model ``c)'', our intention in the present work is to take a step backwards and look at the problem from a more general point of view before embarking in the methodical study of a particular type of model. \\\\ \\\\ The article is organized as follows: in Section II, we analyze the generic form of the rotation curves of galaxies in a general relativistic context. In section III, we comment on the Newtonian approximation and on the embedding of the galaxy in the large scale space-time. In Section IV, we discuss the additional information that can be obtained about the metric from other considerations, specifically the deflection of light by the galaxy. Section V reviews the spontaneous scalarization scenarios. In Section VI, we review the non-minimally coupled global monopole model and discuss its shortcomings. Finally, in Section VII we offer a discussion and analyze the directions for further developments. \\bigskip ", "conclusions": "The galactic rotation curves continue to pose a challenge to present day physics as one would want to understand not only the nature of the dark matter that is associated with them but also the reason behind their universality (i.e., why is it distributed within a galaxy in a way that leads to almost flat rotation curves ?, and why is the amount of dark matter present in a galaxy so well correlated with the luminous matter ? \\cite{persic,tully}). Models based in ordinary physical objects could already be facing problems (depending on the exact value of the Hubble constant \\cite{copi}) in view of the bounds that big bang nucleosynthesis impose on the baryon content of the universe. Models based on particle physics are the most commonly considered (usually within a Newtonian scheme) but they need to address the nature and the distribution problems separately, leading to a larger number of hypothesis and surprinsing coincidences \\cite{sellwood}. In view of the recent cosmological measurements and the theories that have been put forward to explain them \\cite{5essence}, one is naturally lead to consider alternative models based in the introduction of long range coherent fields \\cite{peebles}. In this work, we have given a review of various types of approaches to these questions indicating in each case the problems and advantages. We have argued that so far the most promising and simple approach would involve global monopoles with some sort of nonminimal coupling to gravity. This remains for the future to establish how far can this sort of ideas be pushed towards the goal of making a realistic and compelling model for the dynamics and evolution of galaxies. In particular any such model must also be studied in the context of cosmological perturbations, large scale structure and the CMB. In this regard we should point out that the simplest models of topological defects as seeds for structure formation seem to be incompatible with the acoustic peak in the CMB anisotropies detected by Boomerang and Maxima \\cite{boommax}. However, all these studies have considered the simplest minimally coupled models and it is unclear how would the models of the type being analyzed here behave in this respect. Finally, we should mention that the currently favored cosmological scenarios require at least two hypothetical components: the Cold Dark Matter (usually in the form of WIMPS) necessary for the structure growth and the dark matter in galaxies and clusters, and the cosmological constant $\\Lambda$ which provides the closure density (as required by inflation) as well as the repulsive component that seems to be required in order to account for the observations of the luminosity-distance of high red shift (type Ia) supernovae \\cite{snI}. The fact that the non-minimally coupled monopoles exhibit both an attractive regime at short distances and a repulsive regime at large distances leads us to speculate whether these type of models can be used to explain the two aspects of the unobserved energy content of the universe in terms of a single hypothetical component. Needless is to say that all these aspects will require intense further exploration, which we hope to undertake in the near future. \\vskip 1cm {\\bf Acknowledgments} \\bigskip U.N. is supported by a CONACyT postdoctoral fellowship grant 990490; M.S. and D.S. acknowledge partial support from DGAPA-UNAM Project No. IN121298 and from CONACyT Projects 32551-E and 32272-E. Authors thank the supercomputing department of DGSCA-UNAM. \\newpage" }, "0011/astro-ph0011524_arXiv.txt": { "abstract": "I give a brief overview of cosmic ray physics, highlighting some key questions and how they will be addressed by new experiments. ", "introduction": "There are several new experiments in cosmic-ray physics and related fields running now or planned for the near future. These include measurements of antiprotons with balloons and from space, new gamma-ray detectors to cover the full energy range from $<100$~MeV to $>10$~TeV, studies of the knee region with a variety of experiments, giant air shower detectors for the highest energy particles, and huge, deep neutrino detectors to find astrophysical sources of high energy neutrinos. In this talk, I will review the key questions driving these experiments and, where possible, put them in a larger context and relate them to each other. After some brief remarks about the sun as a cosmic-ray source, I review cosmic rays of galactic origin and then extragalactic cosmic rays. Energy content of the various components of the cosmic radiation and the corresponding power to maintain the observed fluxes are used as a guide to possible sources. ", "conclusions": "" }, "0011/astro-ph0011238_arXiv.txt": { "abstract": "We present a technique that combines Zeeman Doppler imaging (ZDI) principles with a potential field mapping prescription in order to gain more information about the surface field topology of rapid rotators. This technique is an improvement on standard ZDI, which can sometimes suffer from the suppression of one vector component due to the effects of stellar inclination, poor phase coverage or lack of flux from dark areas on the surface. Defining a relationship beween the different vector components allows information from one component to compensate for reduced information in another. We present simulations demonstrating the capability of this technique and discuss its prospects. ", "introduction": "The solar-stellar analogy is commonly invoked to explain stellar activity phenomena in terms of solar features such as prominences and flares \\cite{radick91}. A solar-type dynamo mechanism is also thought to operate in other late-type stars. However it is unclear if this exact same mechanism will operate in stars covering a wide range of convection zone depths, rotation rates and masses. Through the Mt Wilson H\\&K survey, which has monitored stellar rotation periods and activity cycles using chromospheric emission variations spanning a period of over 30 years, we know that many cool stars also display regular solar-type activity cycles. However, other stars in the survey show irregular variations and some show no cyclic patterns over the time span of the observations \\cite{donahue96,baliunas95}. \\scite{saar99} and \\scite{brandenburg98} find that the type of dynamo activity appears to change with rotation rate and age in a study using data from the Mt Wilson survey as well as other photometric studies of cool stars. Doppler imaging techniques can map $T_{\\rm eff}$ flux distributions on the surfaces of rapidly rotating cool stars \\cite{cameron92doppler,piskunov90,rice89,vogt88}. Starspot maps of cool stars taken a few rotation cycles apart allow us to determine the differential rotation rate on rapid rotators accurately \\cite{donati97doppler,barnes2000pztel,petit2000HR1099,donati2000RXJ}. As \\scite{cameron00omega} report, these measurements lend support to differential rotation models in which rapid rotators are found to have a solar-type differential rotation pattern \\cite{kitchatinov99}. Doppler maps of rapidly rotating stars show flux patterns which are very different to those seen on the Sun, with polar and/or high-latitude structure often co-existing with low-latitude flux \\cite{deluca97,strassmeier96table}. This is hard to reconcile with the solar distribution where sunspots and therefore the greatest concentrations of flux tend to be confined between $\\pm 30^{\\circ}$ latitude. Does this indicate different dynamo modes are excited in these stars ? Models by \\scite{granzer00} and \\scite{schussler96buoy} explain the presence of mid-to-high latitude structure on young, rapidly rotating late-type stars in terms of the combined effects of increased Coriolis forces and deeper convection zones working to ``pull'' the flux closer to the poles. Even though \\scite{granzer00} find that both equatorial and polar flux can exist on T Tauri stars that have very deep convection zones, in terms of flux ``slipping'' over the pole, they find the emergence of low-latitude flux on K dwarf surfaces difficult to explain. The technique of Zeeman Doppler imaging (ZDI) allows us to map the magnetic field distributions on the surfaces of rapid rotators using high resolution circularly polarized spectra \\cite{semel89}. Magnetic field maps of the subgiant component of the RS CVn binary, HR1099 ($P_{rot}=2.8$d), and the K0 dwarfs, AB Dor ($P_{rot}=0.5$d), LQ Hya ($P_{rot}=1.6$d) have been presented in several papers \\cite{donati97doppler,donati99doppler,donati99lq}. These maps all show patterns for which there is no solar counterpart: strong radial and azimuthal flux covers all visible latitudes. On AB Dor, strong unidirectional azimuthal field encircling the pole is consistently recovered over a period of three years. This is very different to what is seen on the Sun, where hardly any azimuthal flux is observed at the surface and mean radial fields are much smaller than those observed in ZD maps. While ZDI is an important tool in measuring the surface flux on stars, some questions about the authenticity of the features in these Zeeman Doppler maps remain. ZDI makes no assumptions about the field at the stellar surface; the radial, azimuthal and meridional vectors are mapped completely independently \\cite{hussain99thesis,donati97recon,brown91zdi}. For the true stellar field, however, the physics of the stellar interior and surface will determine the relationship between the field components. Additionally, poor phase coverage is found to lead to an increased amount of cross-talk between radial, meridional and azimuthal field components in ZDI maps \\cite{donati97recon}. Due to the presence of starspots, these maps are also flux-censored. Indeed \\scite{donati99doppler} and \\scite{donati97doppler} find that the low surface brightness regions on AB~Dor appear to correlate with the strongest regions of radial magnetic flux. For these reasons Zeeman Doppler (ZD) maps may not present an accurate picture of the surface field on these stars. By mapping potential fields on the surfaces of stars, we can evaluate more physically realistic models of the surface flux distribution. While we do not necessarily expect the field to be potential locally, this assumption should be adequate on a global scale \\cite{demoulin93}. \\scite{jardine99pot} evaluate the consistency of Zeeman Doppler maps for AB Dor with a potential field using maps obtained over three years. Although they find that the potential field cannot reproduce features such as the unidirectional azimuthal flux found at high latitudes, they do find a good correlation between the ZD azimuthal map and potential field configurations produced by extrapolating from the ZD radial map. The technique presented in this paper reconstructs potential field distributions directly from observed circularly polarized profiles. This technique allows us to produce different configurations which are more physically realistic and which provide more information about the surface topology. ", "conclusions": "Many limitations in ZDI arise from the sensitivity of circularly polarized profiles (Stokes V parameter) to the line-of-sight components of magnetic fields. However, until spectropolarimetric instrumentation is sufficiently advanced so that all four Stokes parameters can be observed simultaneously for cool stars this problem is unlikely to be resolved. Currently, the four Stokes parameters can only be measured in bright, highly magnetic A and B stars \\cite{wade00stokes,wade00ap}. The lack of physical realism in ZD maps also allows for unrealistic magnetic field distributions. By limiting ZD reconstructions to potential field distributions we can overcome some of these problems. In mapping potential fields, we introduce a mutual dependence on all three surface field vectors thereby recovering more field information. The benefits of this method are threefold: firstly, in regions where one vector is suppressed due to the inclination of the star (such as regions of low-latitude meridional field) there is still a reasonably high contribution from the other two and thus field information is reconstructed more reliably; secondly, in areas of bad phase coverage, azimuthal fields can sometimes still be recovered (as their signatures are strongest about $45^{\\circ}$ away from the previous point of observation) and the mutual dependence of the field vectors can be exploited to recover some radial and meridional field information; and finally, in regions such as starspots where the field is thought to be mainly radial but where there is little flux contribution from all three vectors, the field information in the surrounding spot areas can be used to recover more of the lost polarized signal. In addition to these improvements, potential field configurations should allow us to test if the unidirectional band of azimuthal field seen in ZD images of AB Dor is necessary to fit the data. Such a band is incompatible with a potential field and its presence challenges our current understanding of magnetic field generation. If datasets are available that span a sufficiently long time-span it will be possible to monitor which modes are consistently active on these stars which will give us insight into the dynamo mechanisms operating in cool stars. Extrapolating these surface potential field models to the corona will enable us to model the coronal topology of the star and thus to predict where stable prominences may form in these stars. Finally, special care must be taken when applying this technique to rotationally broadened Stokes V spectra from cool dwarfs. The weighting scheme used will have to be tested more thoroughly in order to ensure that we are not sacrificing realism for the ``simplest images'' (in terms of information content). More realistic images may be obtained by incorporating a form of entropy which pushes the field distribution to a power law form, a possibility we are investigating further." }, "0011/astro-ph0011074_arXiv.txt": { "abstract": "The Red-Sequence Cluster Survey (RCS) is a 100 deg$^2$ optical survey for high-redshift galaxy clusters. One of the goals of the survey is a measurement of $\\Omega_m$ and $\\sigma_8$ via the evolution of the mass spectrum of galaxy clusters. Herein we briefly describe how this will initially be done, and also demonstrate the eventual power of the RCS for this type of measurement by a qualitative analysis of the first 1/10th of the survey data. ", "introduction": " ", "conclusions": "" }, "0011/astro-ph0011242_arXiv.txt": { "abstract": "{ The narrow-line Seyfert 1 galaxy \\mrk\\ was observed for 60 ks with the \\xmm\\ observatory. The source shows a complex X-ray spectrum. The 2-10 keV spectrum can be adequately represented by a power law and broad Fe K$\\alpha$ emission. Between 0.7 and 2 keV the spectrum is harder and exhibits a flux deficit with respect to the extrapolated medium energy slope. Below 0.7 keV, however, there is a strong excess of emission. The RGS spectrum shows an edge-like feature at 0.7 keV; the energy of this feature is inconsistent with that expected for an OVII edge from a warm absorber. \\mrk\\ varies by a factor of $\\sim$ 2 in overall count rate in the EPIC and RGS instruments on a timescale of a few thousand seconds, while no significant flux changes are observed in the ultraviolet with the OM. The X--ray variability is spectrally dependent with the largest amplitude variability occurring in the 0.4-2 keV band. The spectral variability can be explained by a change in flux and slope of the medium energy continuum emission, superimposed on a less variable (or constant) low energy emission component. ", "introduction": "\\label{sec:introduction} X--ray emission currently provides the best tracer of conditions close to the supermassive black hole central engines of AGN. However the origin of AGN X--ray emission remains poorly understood. Between 1 and 20 keV AGN generally have a power law spectrum modified by the effects of reflection and absorption (Pounds \\etal \\cite{pounds90}); below 1 keV many AGN also show some `soft excess' emission (Walter \\& Fink \\cite{walter93}). The majority of recent models for AGN X--ray emission explain the soft excess as emission from an optically thick accretion disk, and the power law component as Compton upscattering of soft photons by a hot corona. \\mrk\\ is a nearby (\\(z=0.0129\\)), soft X--ray bright (0.5 - 2 keV flux of $\\sim 10^{-11}$ erg s$^{-1}$ cm$^{-2}$) and variable (factor 2 in \\(\\sim\\) 1000s, Leighly \\etal \\cite{leighly96}) narrow line Seyfert 1 galaxy observed through a relatively small Galactic column of \\(\\sim 1.8 \\times 10^{20}\\) cm\\(^{-2}\\). It is therefore an extremely good candidate for studying the dynamical and radiative processes which give rise to the X--ray emission in AGN. In this letter we present the spectrum and spectral variability of \\mrk\\ as observed by the combination of instruments on board \\xmm\\ during the performance verification (PV) observation. ", "conclusions": "We have presented RGS, EPIC and OM spectra and lightcurves of \\mrk\\ from the 60 ks \\xmm\\ PV phase observation. In all three EPIC cameras the spectra above 2 keV are well described by a power law with a broad Fe K$\\alpha$ emission line. Relative to the power law, the EPIC spectra show a deficit of photons between $\\sim$ 0.7 and 2 keV and excess emission at lower energies. The RGS spectra contain no strong, narrow emission lines. Warm absorber edges are not seen at the expected energies of OVII and OVIII in the RGS spectra; an edge-like feature at 0.7 keV could only be due to OVII if the absorbing material were redshifted more than 10,000 km/s relative to the system. Significant variability is detected in the X--ray lightcurves and hardness ratios, while no variability is detected in the UV with the OM. A variable warm absorber is not required to account for the spectral variability, which can be explained by a medium energy continuum component which softens as it increases in flux, with an additional contribution from a less variable emission component at low energy. Changes in magnetic fields threading the accretion disk corona, or the spectrum of particles injected into the corona, can explain the observed spectral variability." }, "0011/astro-ph0011418_arXiv.txt": { "abstract": "Rossi X-ray Timing Explorer All Sky Monitor observations of \\src\\ show that the source experienced an outburst in January to April 2000 reaching a peak luminosity of greater than \\sqig10$^{38}$ ergs s$^{-1}$. RXTE Proportional Counter Array observations during this outburst reveal the presence of pulsations with a 2.37s period. However, optical photometry of the optical counterpart showed the source to be still significantly fainter than it was more than half a year after the outburst in the 1970s when \\src\\ was discovered. ", "introduction": "The first three X-ray point sources to be discovered in the Small Magellanic Cloud (SMC) were SMC X-1, SMC X-2 and SMC X-3, all of which are thought to be X-ray binaries. While SMC X-1 is a persistent pulsing source (e.g. Wojdowski et al. 1998 and references therein), SMC X-2 and X-3 are transient sources that were discovered in 1977 with SAS-3 (Clark et al. 1978) which were both initially very bright at a level of \\sqig 1.0 $\\times$ 10$^{38}$ ergs s$^{-1}$ (2 $-$ 11 keV) for assumed distances of 71 kpc and remained in outburst for approximately one month (Clark, Li, \\& van Paradijs 1979). Both sources were also detected with HEAO-1 A-2 experiment (Marshall et al. 1979). However, they have since exhibited little activity. SMC X-2 was not detected in an Einstein IPC survey (Seward \\& Mitchell 1980) and was seen in only one of two observations made with the ROSAT PSPC detector (Kahabka \\& Pietsch 1996, Haberl \\& Sasaki 2000) at a luminosity of 2.7 $\\times$ 10$^{37}$ ergs s$^{-1}$ (0.15 $-$ 2.4 keV). SMC X-3 was also not detected by Seward \\& Mitchell (1980) but was seen in one ROSAT HRI observation (Haberl \\& Sasaki 2000). An optical counterpart for SMC X-2 was proposed by Crampton, Hutchings, \\& Cowley (1978) which was resolved into a pair of stars, ``A'' and ``B'', with the fainter object (``B''), a main-sequence Be star, identified as the likely counterpart by Murdin, Morton, \\& Thomas (1979). Infra-red observations of SMC X-2 during a quiescent period are reported by Coe et al. (1997). In none of the previous X-ray observations of SMC X-2 or SMC X-3 were any pulsations detected although they are expected to be present as most Be star X-ray sources are X-ray pulsars. Another bright X-ray pulsar candidate within the SMC is the little studied source H0107-750 which also has a Be star as its proposed optical counterpart (Whitlock \\& Lochner 1994 and references therein). Although no other SMC X-ray pulsars apart from SMC X-1 were known for many years there has recently been an explosion in the discovery of SMC X-ray pulsars. Hughes (1994) discovered a 2.7s pulsar, and three pulsars were found to be active in a single RXTE observation of the vicinity of SMC X-3 (Corbet et al. 1998). Further RXTE, ASCA and SAX observations have now lead to the identification of over 20 X-ray pulsars in the SMC (e.g. Haberl \\& Sasaki 2000, Yokogawa et al. 2000, and references therein) and their variability and several optical identifications suggests that at least the vast majority are transient Be star systems. The number of known Be star X-ray binaries in the SMC is thus now starting to rival the number known in the Galaxy (e.g. Bildsten et al. 1997). In this paper we report the detection of activity from the vicinity of SMC X-2 in early 2000 with the Rossi X-ray Timing Explorer All-Sky Monitor (ASM). Subsequent observations with the RXTE Proportional Counter Array (PCA) then revealed the presence of 2.37s pulsations. Contemporaneous optical photometry of the optical counterpart are compared to the initial discovery photometry. Note that in this paper we use an assumed distance of 65 kpc to the SMC for consistency with several recent publications (e.g. Kahabka \\& Pietsch 1996, Haberl \\& Sasaki 2000) which is somewhat smaller than the 71 kpc adopted by Clark et al. (1978). ", "conclusions": "Given the large number of X-ray pulsars in the SMC can we be certain that the new pulsar that we have detected is indeed SMC X-2? The pulse period is rather short and we can make use of the result of Stella, White \\& Rosner (1986) who noted an anti-correlation between pulse period and maximum X-ray luminosity. SMC X-2 at its peak observed with SAS-3 and the RXTE ASM was rather bright at $>$ \\sqig 10$^{38}$ ergs s$^{-1}$ which, together with the pulse period that we measure, would be consistent with the relationship found by Stella et al. The vicinity of SMC X-2 may also be less densely populated with active pulsars than the region around SMC X-3. After emission had been detected from the vicinity of SMC X-3 with RXTE subsequent analysis revealed the presence of three pulsars (Corbet et al. 1998), while ASCA observations showed that none of these were in fact SMC X-3 itself. However, in the case of SMC X-2, we only detect one pulsar and, in addition, its location is consistent at the \\sqig 3 arc minute level with SMC X-2. The brightening observed in the ASM light curve for SMC X-2 is also consistent with the source being not more than a few arcminutes from SMC X-2. Further evidence still for the identification of the 2.37s pulsar with SMC X-2 comes from the detection of this source by Torii et al. (2000) with the Gas Imaging Spectrometer experiments on board ASCA. The GIS is an imaging instrument which gives source locations with 90\\% errors of 1 to 2 arc minutes radius (Tashiro et al. 1995). In the ASCA GIS survey of the SMC (Yokogawa et al. 2000) 12 X-ray pulsars and 8 candidate X-ray pulsars were identified. We thus obtain an approximate estimate of the chance of finding any X-ray pulsar or candidate within 2 arc minutes of SMC X-2 in a snap-shot survey of a few percent.\\ Another independent estimate of the probability that the source that we have observed is SMC X-2 comes from the serendipitous RXTE slew detection (observation \\#1). The slews across the SMC that occur during routine RXTE observations provide a monitor for bright new sources within this entire region. There are currently only five sources known in the SMC with luminosities \\sqig10$^{38}$ ergs s$^{-1}$: SMC X-1, SMC X-2, SMC X-3, H0107-750, and the 31 second pulsar XTE J0111.2-7317 (Chakrabarty et al. 1998). The probability of finding a new source with a comparable brightness at a position consistent with one of these known bright sources within the approximately 2\\degrees\\ $\\times$ 2\\degrees\\ extent of the SMC is less than 1\\%. Together, the comparable luminosity observed with RXTE and SAS-3 and the positional coincidence at the few arc minute level, strongly suggest that the source we have observed is indeed SMC X-2. The evidence would be strengthened further still if the position of the 2.37 s pulsar could be localized more precisely or evidence found for pulsations at this period in the archival SAS-3 or HEAO-1 observations. The outburst we have observed with RXTE could be due to either changes in the circumstellar envelope of the Be star, associated with orbital variability, or both of these effects. We note that no other outburst of a similar size is apparent in the ASM light curve which would argue for a short lived change in the circumstellar envelope. If \\src\\ has properties consistent with the general correlation between orbital and pulse periods exhibited by most Be/neutron star binaries (Corbet 1986) then an orbital period of order 20 days to within a factor \\sqig2 might be expected. However, no evidence for any periodic variability is found in the ASM light curve. During the course of our observations the pulse period of \\src\\ increased and was longer still at 2.37230 $\\pm$ 0.00004 s when ASCA observed on April 25-26 (Torii et al. 2000). This change in pulse period may be due to orbital motion and would correspond to a velocity change of \\sqig 100 km s$^{-1}$. While there are currently too few pulse period measurements to place any strong constraints on the orbital parameters, if \\src\\ returns again to an active state it is possible that additional observations may yield the orbital period, either from monitoring the pulse period or from periodic outbursts that the ASM could observe. The pulse period that we find for \\src\\ is among the shortest currently known for high mass X-ray binaries (e.g. Bildsten et al. 1997) especially for those systems which contain Be star primaries. As noted above, this short pulse period is consistent with the high source luminosity. SMC X-3, which also exhibited a high maximum luminosity, may also be expected to have a relatively short pulse period. Previously Clark et al. (1978, 1979) argued that the high luminosities of SMC X-1, X-2 and X-3 implied that the luminosity distribution of the SMC X-ray sources is shifted toward higher luminosities compared to Galactic systems. However, now that the transient X-ray pulsar population of the SMC is known to be so large this shift to higher luminosities is called into question (see e.g. Haberl \\& Sasaki 2000) as the earlier observations were only detecting the high-luminosity tip of the large SMC X-ray pulsar population. Extensive studies of these objects down to low flux levels are required to fully establish their overall properties as a class." }, "0011/astro-ph0011132_arXiv.txt": { "abstract": "I summarize some of the consequences for the optical and very-near-infrared spectra of T dwarfs (in particular) and brown dwarfs (in general) of their possible dominance by the neutral alkali metal lines. As a byproduct of this study, I estimate the true optical color of ``brown'' dwarfs. ", "introduction": "The early discovery phase for L dwarfs and T dwarfs has ended and a major focus is now on their characterization. The atmospheres of brown dwarfs are dominated by H$_2$, H$_2$O, CH$_4$, NH$_3$, the neutral alkali metals, and grains, but how theory translates this basic knowledge into effective temperatures, gravities, and compositions has yet to be determined. Establishing the spectral and color diagnostics that are most appropriate for L/T studies is complicated by ambiguities in the cloud/grain models and a paucity of opacity data. In particular, though T dwarfs are being informally defined by their methane features at 1.7 \\mic and 2.2 \\mic, the methane database itself is far from complete. The methane opacities on the red side of the $H$ band are certainly in error by a factor of 3 to 5 (witness Gliese 229B\\cite{leg99}) and the hot bands are completely missing. The latter means that even the sign of the opacity's dependence upon temperature can be in error. Nevertheless, there has been great overall progress towards understanding what makes these objects unique and what their spectra are telling us. In this paper, I sidestep a comprehensive study of these issues and summarize three interesting topics in brown dwarf theory that have emerged of late. They are 1) what determines T dwarf spectra shortward of 1.0 micron, 2) what is the true color of a ``brown\" dwarf, and 3) what is the effect of heavy element depletion (``rainout'') on the abundance profiles of the neutral alkali metal atoms. A subtext of this contribution is the central importance of the alkali metals in spectrum formation. ", "conclusions": "L and T dwarf spectra are unique among ``stars\" and require new databases, approaches, and thinking to fully understand. Exploring as we are new worlds, we will require new tools and instincts with which to navigate. Along with accurate cloud models, methane, and water, the alkali metals hold the key to unraveling the mysteries of the substellar objects that we now know inhabit the solar neighborhood in abundance. \\subsubsection{Acknowledgments:\\\\} I thank my long-time collaborators, Jonathan Lunine, Bill Hubbard, and Mark Marley, for simulating input and both Davy Kirkpatrick and Neill Reid for an advanced glimpse at their stunning 2MASSW J1507 spectrum. This work was supported in part by NASA under grants NAG5-7073 and NAG5-7499." }, "0011/astro-ph0011060_arXiv.txt": { "abstract": "We probe the region around the protostar HH108MMS by deep mid infrared photometric and polarimetric imaging. The protostar is detected at 14$\\mu$m in absorption against the diffuse background. Next to HH108MMS, we find a second absorbing core, named Q1, and the young stellar object IRAS18331--0035 which is more advanced in its evolution and already seen in emission at 12$\\mu$m and 14$\\mu$m. HH108MMS, Q1 and IRAS18331--0035 form a triplet along an extended filamentary absorption feature. From the variation of the surface brightness across the source, we derive for HH108MMS and Q1 the optical depth and density profile. Along the axes which are parallel to the filament, the density distributions follow a $\\rho\\propto r^{-1.8}$ power law. We estimate that the intensity of the background radiation at 14$\\mu$m is about two times stronger than the intensity of the interstellar radiation field in the solar neighborhood. The present photometric data of IRAS18331--0035 between 12$\\mu$m and 1.3mm can be explained by a central source with a luminosity of 2.5\\,\\Lsun\\ that is surrounded by a spherical cloud of 1.1\\,\\Msun\\ with a $1/r$ density distribution. As HH108MMS is also seen in the millimeter dust emission, we can derive the ratio of the dust extinction coefficients at 14$\\mu$m and 1.3mm and obtain $\\kappa_{14\\mu\\rm m} / \\kappa_{1300\\mu \\rm m}\\sim 470$. Because models for the dust in the diffuse interstellar medium predict a ratio of around 2000, our value points to fluffy composite grains which are expected to prevail in dense and cold environments. \\noindent First mid infrared polarisation images of pre--stellar absorbing cores are presented. At 12$\\mu \\rm{m}$ and 14$\\mu \\rm{m}$ the polarisation of the region around HH108MMS is strong ($\\geq 15\\%$) and tightly correlated with the source triplet. We demonstrate that the high degree of polarisation can be explained by extinction of rotationally aligned dust particles of moderate elongation. ", "introduction": "\\noindent Herbig Haro objects mark shock regions in the outflow from very young stars. The 1.3mm continuum survey of Herbig Haro objects by Reipurth et al.~(1993) indicates that the young stars, which are the energy sources of the outflow, are surrounded by dusty envelopes of about 0.1 -- 3 solar masses. The circumstellar material is so massive that the stars are probably still in the accretion phase. In a follow--up study of the 1.3mm dust continuum, Chini et al. (1997) discovered in the Serpens star forming region which is at a distance $D= 310$\\,pc (De Lara et al. 1991) a protostellar candidate located 71$''$ (0.11pc) away from the source IRAS18331--0035. The latter is believed to be the driving engine of the two aligned nearby bow shocks HH108 and HH109 (Reipurth \\& Eiora 1992, Ziener \\& Eisl\\\"offel 1999); they lie to the South-West of IRAS18331--0035, about 0.1pc and 0.21pc away. In this paper, we present photometric and polarimetric mid infrared images of the region. ", "conclusions": "\\noindent Among many other models, the self--similarity solutions of Shu et al.~(1977) for the early evolution of protostars predict in the absence of rotation and magnetic fields a $\\rho \\propto r^{-2}$ profile in the outer envelope and a less steep distribution $\\rho \\propto r^{-3/2}$ in the inner part. Basu \\& Mouschovias (1995), who include magnetic fields and rotation, also find a power law density distribution in the envelope with an exponent between $-1.5$ and $-1.8$ and a constant density central region of $\\sim$50\\,AU size. Masunaga \\& Inutsuka (1999) perform calculations for the early collapse (before dissociation of H$_2$ in the core) of a 1\\,M$_\\odot$ star. The luminosity stays below 0.1\\,L$_\\odot$, there is a core of 5\\,AU and the density goes like $r^{-2}$ in the envelope. Their models should be applicable to HH108MMS and Q1 of the present paper which also have no mid--IR embedded source. \\noindent In a few cases, the predicted theoretical density profiles could be checked by 1.3mm observations of the dust emission (Ward--Thompson et al.~1994, Andr\\'e et al.~1996, Ward--Thompson et al.~1999) and were generally corroborated. In this method, the conversion of fluxes into absolute column densities depends on the dust absorption coefficient at 1.3\\,mm and the grain temperature, whereas the power law exponent of the density distribution is sensitive to the temperature gradient. \\noindent On the other hand, our derivation of the optical depth and the ensuing density profile from absorption measurements is temperature independent. ISOCAM observations similar to ours were also carried out by Bacmann et al.~(2000). For the sources which they present, the maximum optical depth at the cloud center is much smaller than those for HH108MMS and Q1, so we probe deeper into the cloud, still the density profiles in the {\\it envelopes} roughly agree with what we derive. In the {\\it cloud center} ($\\le 2000$AU), they find a flattening of the column density which we do not see. Such a flattening is expected to occur in the central 1000AU of an isothermal sphere with temperature $T \\sim 10$K, radius of 10000AU and mass of 1\\Msun\\ simply by solving the hydrostatic equation (Bonnor 1956). \\noindent \\noindent The measured mid infrared to millimeter dust extinction ratio of $ \\kappa_{14\\mu \\rm{m}} / \\kappa_{1300\\mu \\rm{m}} \\sim 470 $ should be compared with dust models of protostellar environments. The ratio is a factor 4 lower than expected for dust in the diffuse medium and it indicates that the grains in the dense and cold environment of the protostar HH108MMS are, as expected (Ossenkopf 1993), of rather fluffy and composite nature. The measured mid infrared to millimeter dust extinction ratio of HH108MMS is already tending towards somewhat elongated grain structures: In the fluffy composite but spherical dust model by Kr\\\"ugel \\& Siebenmorgen (1994) a ratio of $\\kappa_{14\\mu\\rm{m}} / \\kappa_{1300\\mu \\rm{m}} \\sim 1000$ is found. Because elongated particles are much better antennas at 1.3mm, they can give a lower value. A proof that the dust in the absorbing cores and the IRAS source is indeed of elongated structure is found by the ISOCAM polarisation measurements. \\noindent As the filament is seen in absorption against the background and assuming that the majority of the grains are still sub-micron sized particles, one may neglect dust scattered light at 14$\\mu$m. Consequently the most plausible mechanism to produce the polarisation is dichroism. For polarisation due to extinction of elongated spinning and aligned dust particles, the polarisation vectors (Figs.~\\ref{IRAS18331_pol_lw3.ps} and \\ref{IRAS18331_pol_lw10.ps}) indicate the magnetic field direction. It appears that the magnetic field vectors are roughly aligned with the absorbing filament. Such ordered fields are also detected from 850$\\mu {\\rm m}$ maps of thermal dust emission of prestellar cores (Ward-Thompson et al.~2000) and FIR polarimetry on the protostar IRAS20503+6006 (Clements et al.~1999)." }, "0011/astro-ph0011583_arXiv.txt": { "abstract": "Gamma-ray bursts provide what is probably one of the messiest of all astrophysical data sets. Burst class properties are indistinct, as overlapping characteristics of individual bursts are convolved with effects of instrumental and sampling biases. Despite these complexities, data mining techniques have allowed new insights to be made about gamma-ray burst data. We demonstrate how data mining techniques have simultaneously allowed us to learn about gamma-ray burst detectors and data collection, cosmological effects in burst data, and properties of burst subclasses. We discuss the exciting future of this field, and the web-based tool we are developing (with support from the NASA AISR Program). We invite others to join us in AI-guided gamma-ray burst classification (http://grb.mnsu.edu/grb/). ", "introduction": "Understanding the physics of a class of astronomical objects depends on identifying intrinsic behaviors. When two or more subclasses are present, each subclass is defined in terms of its own intrinsic behaviors. The process of identifying behavioral characteristics is difficult when the objects' observed characteristics (or {\\em attributes}) overlap. Such is the case for cosmic gamma-ray bursts (GRBs), which have a large spread in observed attribute values. Some GRB attribute dispersion is intrinsic, some is caused by measurement error, some is due to systematic ({\\em e.g.} instrumental and sampling) biases, and some is caused by the presence of subclasses. GRB subclass behaviors are difficult to delineate from other causes of attribute dispersion. Two GRB subclasses have been known to exist for some time \\cite{C74} \\cite{K93}, but it has been difficult to assign individual GRBs to a class because of attribute overlap. Class assignment has been complicated even more by the statistical clustering identification of a third GRB subclass \\cite{M98}; properties of this third subclass overlap those of the other two. GRB classification can be aided by Knowledge Discovery in Databases (KDD) \\cite{D00}. The approach uses pattern recognition algorithms from the Artificial Intelligence (AI) branch of computer science to find behaviors indicative of subclasses. KDD offers a methodology by which meaningful information can be extracted from large volumes of data. The KDD process (Figure \\ref{fig:fig0}) is composed of data pre-processing and storage (data warehousing), data mining (clustering software), and scientific/logical assessment. Statistical and systematic effects (e.g. instrumentation and sampling biases) can be identified and even removed in the assessment step. \\begin{figure}[htbp] \\begin{center}\\small \\unitlength=0.66mm \\linethickness{0.4pt} \\begin{picture}(245,80) \\put(5,20){\\dashbox{1}(45,50)[ct] {\\shortstack{\\\\ \\\\ Data Warehousing}}} \\put(12.5,30){\\framebox(30,30)[cc] {\\shortstack{Pre- \\\\ Processed \\\\ GRB Data}}} \\put(42.5,42){\\line(1,0){17}} \\put(42.5,48){\\line(1,0){17}} \\put(62.5,45){\\line(-1,+1){6}} \\put(62.5,45){\\line(-1,-1){6}} \\put(55,20){\\dashbox{1}(45,50)[ct] {\\shortstack{\\\\ \\\\ Data Mining}}} \\put(62.5,30){\\framebox(30,30)[cc] {\\shortstack{Classifier \\\\ Shell}}} \\put(93,42){\\line(1,0){8.5}} \\put(93,48){\\line(1,0){8.5}} \\put(104.5,45){\\line(-1,+1){6}} \\put(104.5,45){\\line(-1,-1){6}} \\put(105,30){\\framebox(30,30)[cc] {\\shortstack{Scientific \\\\ and \\\\ Logical \\\\ Assessment}}} \\put(135.5,42){\\line(1,0){5.5}} \\put(135.5,48){\\line(1,0){5.5}} \\put(144.5,45){\\line(-1,+1){6}} \\put(144.5,45){\\line(-1,-1){6}} \\put(120,30){\\line(0,-1){15}} \\put(49.5,14){Iteration} \\put(120,15){\\line(-1,0){49}} \\put(47.5,15){\\line(-1,0){20}} \\put(27.5,15){\\vector(0,+1){15}} \\put(145,30){\\framebox(30,30)[cc] {\\shortstack{GRB \\\\ Subclasses}}} \\put(0,10){\\dashbox{0.5}(140,70)[ct] {\\shortstack{\\\\ \\\\ Gamma-Ray Burst Classification Tool}}} \\end{picture} \\end{center} \\vspace*{-10mm} \\caption{Gamma-Ray Burst Classification Process}\\label{fig:fig0} \\end{figure} AI classifiers are typically {\\em supervised} or {\\em unsupervised}. Supervised classifiers require training {\\em instances} (data elements) in order to develop classification rules for unknown instances. Unsupervised classifiers try to subclassify a data set by searching for clusters in multidimensional attribute spaces. We are developing a web-based tool \\cite{Hag00} for the classification of GRB data (http://grb.mnsu.edu/grb/). The tool contains a preprocessed GRB database, AI classifiers, and data visualization software. In this manuscript we describe some of our initial scientific results concerning GRB data mining with this tool. Additional results have been published elsewhere \\cite{H00}, \\cite{R00}, \\cite{Hak00}. ", "conclusions": "We have demonstrated that data mining techniques can aid the interpretation of scientific data, even with complex and ambiguous GRB data. Data mining demonstrates that some Class 1 (Long) GRBs can develop Class 3 (Intermediate) characteristics via a combination of the hardness intensity relation and the fluence duration bias. Class 3 (Intermediate) GRBs do not appear to represent a separate source population, although they cluster in the duration, fluence, hardness, attribute space. Class 2 (Short) GRBs do appear to represent a separate source population." }, "0011/astro-ph0011530_arXiv.txt": { "abstract": "We study spectroscopically determined iron abundances of $642$ solar-type stars to search for the signature of accreted iron-rich material. We find that the metallicity [Fe/H] of a subset of $466$ main sequence stars, when plotted as a function of stellar mass, mimics the pattern seen in lithium abundances in open clusters. Using Monte Carlo models we find that, on average, these stars have accreted $\\sim0.4M_\\oplus$ of iron while on the main sequence. A much smaller sample of $19$ stars in the Hertzsprung gap, which are slightly evolved and whose convection zones are significantly more massive, have lower average [Fe/H], and their metallicity shows no clear variation with stellar mass. These findings suggest that terrestrial-type material is common around solar type stars. ", "introduction": "The discovery that at least $6-8\\%$ of solar type stars harbor Jupiter-mass or larger bodies, often in small, eccentric orbits \\citep{mq, marcy, butler, mcm} shows that planets are not exceptionally rare. The transiting planet orbiting HD 209458 has a mass of $0.69$ Jupiter masses, and a radius about 1.4 times that of Jupiter. Clearly it is a gas giant like Jupiter or Saturn; the minimum masses of the other known objects suggest that they are also gas giants. The Doppler technique used for most of the discoveries cannot find terrestrial mass objects in AU scale orbits. What fraction of solar type stars have terrestrial-type planets? How could an observer with our current technology located on a star in the solar neighborhood decide if there were any terrestrial-type bodies orbiting the sun? More generally, how could such an observer estimate the fraction of solar type stars having companions of terrestrial (as opposed to gas giant) compositions? In this paper we explore one possible way of addressing the latter question, stellar-mass dependent photospheric metallicity trends. In section \\ref{pollution} we argue that a few Earth masses of rocky/icy material have accreted onto the Sun over its lifetime. In section \\ref{evidence} we discuss one possible observational tracer of similar accretion occurring on other solar type stars. In section \\ref{observations} we examine a sample of stars with spectroscopically measured photospheric iron abundances \\be % {\\rm [Fe/H]}\\equiv\\log\\left[{f_{Fe}\\over f_{Fe,\\odot}}\\right], \\ee % where $f_{Fe}$ is the mass abundance of iron in the photosphere of the star, and $f_{Fe,\\odot}\\approx 1.3\\times10^{-3}$ is the the mass abundance of iron in the photosphere of the sun. We compare the observations to Monte Carlo models of stellar pollution in section \\ref{gamble}. We discuss the implications of our results in section \\ref{discussion} and present our conclusion in the final section. ", "conclusions": "We analyzed some 642 stars from the Cayrel de Strobel catalogue having spectroscopically determined metallicities, and well determined HIPPARCOS parallaxes. Using a large grid of stellar models, we determined the age and mass of the stars in our sample. We then examined the variation of metallicity with stellar age and mass. We have found striking mass-dependent variations in photospheric iron abundances of main sequence solar mass ($0.8-1.8M_\\odot$) stars in the solar neighborhood. These variations mimic the variations seen in lithium abundances. With somewhat less confidence, because of the small sample size, we find that the iron abundances of Hertzsprung-gap stars are on average lower than those of main sequence stars, and that the metallicities of these slightly evolved stars have no mass dependence. These results are consistent with the accretion of an average of $\\sim0.4M_\\oplus$ of iron onto the surface of main sequence stars. This strongly suggests that terrestrial-type material is common around solar type stars in the solar neighborhood." }, "0011/astro-ph0011195_arXiv.txt": { "abstract": "Observations of the cosmic microwave sky are revealing the primordial non-uniformities from which all structure in the Universe grew. The only known physical mechanism for generating the inhomogeneities we see involves the amplification of quantum fluctuations during a period of inflation. Developing the theory further will require progress in quantum cosmology, connecting inflation to a theory of the initial state of the Universe. I discuss recent work within the framework of the Euclidean no boundary proposal, specifically classical instanton solutions and the computation of fluctuations around them. Within this framework, and for a generic inflationary theory, it appears that an additional anthropic constraint is required to explain the observed Universe. I outline an attempt to impose such a constraint in a precise mathematical manner. ", "introduction": "The maps of the cosmic microwave sky provided by recent experiments \\cite{1} signal an important breakthrough in cosmology. The structure revealed provides our most powerful probe of the structure in the Universe at early times, and of its current geometry. The data is consistent the Universe being nearly spatially flat, with a scale invariant spectrum of Gaussian-distributed irregularities in which the overall density fluctuates but the equation of state does not. This simplest form was anticipated on phenomenological grounds as far back as the sixties \\cite{2}, but later emerged as the prediction of simple models of inflation \\cite{3,4,5,6}. The detailed agreement between theory and observations lends impressive support to the view that the Universe is, in some respects at least, remarkably simple. In inflationary theories, inhomogeneities of the required form arise a side-effect of the exponential expansion invoked to explain the size and flatness of the Universe. Quantum mechanical vacuum fluctuations in the inflaton field are stretched to large scales during inflation \\cite{5,6}, later to re-enter the Hubble radius and seed large scale structure. This magnificent mechanism links quantum mechanics and gravity to the observed structure of the Universe. It provides a direct observational probe of quantum gravity, albeit at leading (Gaussian) order. The underlying theory is however only provisional, because we still have no compelling model of inflation and because a complete theory of quantum gravity is lacking. Nevertheless the success of the inflationary explanation of the structure in the cosmic microwave sky forces us to take very seriously the idea that quantum mechanics governed the formation of the Universe we see. This moves the topic of quantum cosmology to centre stage. It is a field that {\\it must} be developed if we are attain a deeper level of understanding in cosmology. The prevalent view is that cosmology will over the next period be observation-driven. It is true that dramatic observational developments are certain, and there will be a lot of phenomenology to do. However if we mean to achieve more than simply measuring the properties of the universe, theoretical developments are also crucial. The great freedom there is in current approaches to phenomenological models of inflation and of the early Universe, combined with the relatively small number of observations (even including the cosmic microwave sky) leads me to conclude that if we are to get anywhere, mathematical requirements of completeness and consistency must play an increasing role. Their power is already convincingly demonstrated by the fact that not one of the existing inflationary models has been made sense of beyond leading (Gaussian) order. In this talk I will concentrate on one particular incompleteness of inflationary theory. The question is why inflation started in the first place. Vilenkin, Linde and Guth argue that the need for a theory of initial conditions is side-stepped because inflation is self-sustaining to the future (i.e. `semi-eternal')\\cite{eternal}. Whilst accepting that some theory of the initial conditions is needed at some level, the claim is that the details of that theory are actually irrelevant, since there would typically be an infinite amount of inflation between ourselves and the beginning, during which details of the initial conditions would have been erased. In this view discussions of the initial conditions prior to inflation are fairly unimportant, since `almost anything' will do. In this talk I criticise the calculational methods which have been used to reach this conclusion, and explain a different point of view according to which inflation is {\\it a priori} very unlikely and in any case only a brief episode in our past. One can easily imagine an infinite number of possible ansatzes for the initial conditions of the Universe. At the present stage of development of theory, many of these might be perfectly consistent with observation. But among the existing proposals, I think the Euclidean no boundary proposal \\cite{hh} is appealing because it is based on simple and general ideas which have a rationale beyond cosmology. In a strong sense I think it is `the most conservative thing you can do'. Of course it may well fail precisely because it is too conservative. Space and time may be emergent rather than fundamental properties. Describing the Universe as a manifold may not be appropriate to its early moments. Nevertheless precisely because the no boundary proposal is {\\it not} just cooked up to make inflation work, its failings and limitations may teach us something deeper about what is in fact required. In this lecture I focus on one particular approach to the no boundary proposal, using a set of classical instanton solutions of the `no boundary' form \\cite{ht,th}. Our work has focussed on using these to compute the the complete Gaussian correlators \\cite{gt,het,hht,ght}, in a generic inflationary model (for related work see \\cite{otha,othb}). These generic instantons are unusual in several respects. In the original `Einstein frame' they exhibit a curvature singularity which in the Lorentzian spacetime is `naked'. Nevertheless the Euclidean action for the solutions is finite, suggesting they should contribute to the Euclidean path integral. Second, there is a one parameter family of solutions, with differing action. This is inconsistent with their being true stationary points of the action, but they may nevertheless be legitimate as `constrained instantons' which are an established tool in other contexts \\cite{affleck}. A second feature is that the quantum fluctuations about these solutions are well defined to leading order, since the Euclidean action selects a unique (Dirichlet) boundary condition at the singularity. This allows one to make unambiguous predictions for the microwave anisotropies in any one of these solutions, which Gratton, Hertog and I have recently computed \\cite{gt,het,ght}. (See also \\cite{hht}). I then describe recent work by Kirklin, Wiseman and myself \\cite{ktw} showing how the singularity apparent in the original `Einstein frame' may actually be removed by a change of field variables. Thus the singularity is really only a coordinate singularity on superspace. One still has the problem that the scalar field potential energy diverges at the singularity, and the scalar field equation remains ill defined there. That problem is overcome if one defines the theory as the limit of a family of theories in which $V(\\phi)$ takes a particular form as $\\phi$ tends to infinity, but approximates the actual potential at arbitrarily large values of $\\phi$. The limit is insensitive to the precise form of the regularised potential at large $\\phi$. The regularised construction allows one to give an improved formulation of the constraint at the singularity, which allows for a detailed computation of the spectrum of homogeneous modes \\cite{gtnew}. The latter, we argue, are actually crucial to a proper interpretion of the instantons. In the regularised theory all quantities appearing are finite. The classical field equations are also satisfied everywhere if one views the instantons as topologically $RP^4$ rather than $S^4$. The same construction is applied to instantons possessing two singularities in the Einstein frame. When the singularities are `blown up' one obtains a regular manifold which is a four dimensional analogue of the Klein bottle. Finally I discuss the problem of negative modes, complicated by the `conformal factor' problem of Euclidean general relativity, due to the non-positive character of the Euclidean action. In the regular variables (and with the regularised potential), the problem of negative modes about singular instantons is well defined, and I briefly mention some of our latest results \\cite{gtnew} indicating that the most interesting singular instantons (describing Universes with a large amount of inflation) do not possess any negative modes, in contrast to the non-singular instantons of Coleman and De Luccia, and Hawking and Moss. I comment on how this observation might help to resolve the `empty Universe problem', if one performs a projection onto Universes containing the observer. ", "conclusions": "" }, "0011/astro-ph0011085_arXiv.txt": { "abstract": " ", "introduction": "Surface photometry is one of the oldest techniques in modern Astronomy. The first attempt of surface photometry was done in Helwan observatory in Egypt by Reynolds 1913. In quanititative analysis of the structure of galaxies, some of the two-dimensional data are reduced to one-dimensional luminosity profiles. The shape of the surface brightness profile took more attention and has been put forward as a new extragalactic distance indicator (c.f. Binggeli \\& Jerjen 1998). NGC 3077 is a member galaxy of the M81 interacting group of galaxies. It lies 46$^{\\prime}$ south-east of the spiral galaxy M81. Two HI bridges connect it with its neighbouring galaxies M81 and M82 (cf. Yun et al. 1994, Thomasson and Donner 1993). It was classified as an Irr II galaxy (Sandage 1961). The centre of NGC 3077 is dusty and exhibits many compact blue knots, in addition to condensed HII regions embedded in diffuse ionized gas (Barbieri et al. 1974). In this paper we performed detailed continuum surface photometry to investigate the morphology and the colour distribution of the galaxy NGC 3077. ", "conclusions": "A study of the brightness and colour distribution of NGC 3077 shows that this galaxy consists of a blue disturbed centre embedded in a smooth elliptical structure; this structural type is common in irregular galaxies. The lack of symmetry in the inner region is due to the presence of a young stellar population reddened by dust. The colours of this young population indicate ages between 4 and 100 Myr, whereas the mean age of the dominant population in the outer region of NGC 3077 is 3$\\cdot 10^9$ yr. The age range of the young central population suggests that the ongoing encounter with M81, which commenced a few hundred million years from now (eg. Thomasson and Donner 1993), serves as a trigger for the star formation in the galaxy's core." }, "0011/astro-ph0011566_arXiv.txt": { "abstract": "We report optical, radio and X--ray observations of a new distant blazar, PMN~J0525-3343, at a redshift of 4.4. The X-ray spectrum measured from ASCA and $Beppo$SAX flattens below a few keV, in a manner similar to the spectra of two other $z>4$ blazars, GB~1428+4217 ($z$=4.72) reported by Boller et al and RXJ~1028.6-0844 ($z$=4.28) by Yuan et al. The spectrum is well fitted by a power-law continuum which is either absorbed or breaks at a few keV. An intrinsic column density corresponding to $2\\times 10^{23}$H-atoms\\ \\psqcm at solar abundance is required by the absorption model. This is however a million times greater than the neutral hydrogen, or dust, column density implied by the optical spectrum, which covers the rest-frame UV emission of the blazar nucleus. We discuss the problems raised and suggest that, unless there is intrinsic flattening in the spectral distribution of the particles/seed photons producing X--rays via inverse Compton scattering, the most plausible solution is a warm absorber close to the active nucleus. ", "introduction": "Active Galactic Nuclei (AGN) at high redshift are powerful tools with which to study the evolution of massive black holes and of their young galaxy hosts. A particularly interesting class is formed by several high redshift ($z>4$), X-ray bright, radio--loud quasars (Fabian et al 1997, 1998; Moran \\& Helfand 1997; Zickgraf et al 1997; Hook \\& McMahon 1998) which present characteristics typical of blazars. One such source, GB~1428+4217, is variable in both the X-ray and radio bands and has a spectral energy distribution which peaks at hard X-ray energies (Fabian et al 1998). Recently this object (Boller et al 2000) and RXJ1028.6-0844 (Yuan et al 2000) have been found to show spectral flattening which is interpreted as due to X-ray absorption, implying an absorbing column density of $\\sim 1.5\\times10^{22}$\\psqcm and $\\sim 2.1\\times 10^{23}$\\psqcm, respectively, if intrinsic. We report here on the discovery and study of a similar object, PMN~J0525-3443 at $z=4.4$, which shows similar spectral flattening. Fiore et al (1998) have earlier found from ROSAT data an apparent systematic decrease with redshift of the spectral slope of the soft X--ray emission of radio-loud quasars from local to $z\\sim 3.9$ objects. The change in spectra with redshift might be associated with an increase with redshift in the amount of absorbing (intrinsic or external) gas. Cappi et al (1997) find evidence for absorption using ASCA data for a small sample of high-redshift ($1.2