{ "0010/astro-ph0010242_arXiv.txt": { "abstract": "The Alpha Magnetic Spectrometer in a precursor version (AMS-01), was flown in June 1998 on a 51.6\\mydeg~orbit and at altitudes ranging between 320 and 390 km, on board of the space shuttle Discovery (flight STS-91). AMS-01 included an Aerogel Threshold \\cher counter (ATC) to separate $\\bar{p}$ from $e^{-}$ and $e^{+}$ from $p$, for momenta below 3.5 \\GeVc. This paper presents a description of the ATC counter and reports on its performances during the flight STS-91. ", "introduction": "The first phase of the AMS experiment (AMS-01) was achieved on board of the space shuttle Discovery, during 10 days in June 1998. The main objective was to test the spectrometer's instrumentation in orbit, in preparation for the second phase that will take place on board of the International Space Station (ISS) for 3 to 5 years. During the shuttle flight, 100 million events were recorded, allowing the fluxes of several particle species ($e^{\\pm}$, $p^{\\pm}$, He) to be measured~\\cite{amsres}.\\\\ The AMS-01 detector included a permanent magnet, a Time-of-Flight scintillation counter (TOF), a silicon tracker (TRK), anti-coincidence scintillation counters (ACC) and an Aerogel Threshold \\cher counter (ATC). A detailed description of the AMS spectrometer may be found in \\cite{amsrep}.\\\\ This paper describes the ATC counter and its performances during the flight on board of Discovery. \\subsection{Role of ATC in AMS-01.} One of the main purposes of the AMS Shuttle flight was to measure cosmic antiproton spectrum for momenta below 3.5~\\GeVc~(the ATC momentum threshold). Antiproton spectrum measurement, as well as positron sprectrum, can be achieved by using the ATC counter :\\\\ $\\bullet$ {\\bf antiprotons :} The major background component to the $\\bar{p}$ sample is expected to come from misidentified electrons. Using the measured electron flux~\\cite{amsres} and the previously measured $\\bar{p}$ flux~\\cite{besspbar}, the signal to background ratio is estimated to be : ${\\bar{p}}/e^{-} \\sim \\!10^{-3}\\!-\\!10^{-2}$ for the considered P range.\\\\ While TOF counters~\\cite{kn:choumilo} (${\\Delta \\beta }/\\beta \\simeq 3.3 \\, \\%$) allow the separation of $\\bar{p}$ and $e^{-}$ below 1-1.5 \\GeVc~(fig.~\\ref{tofatc}), ATC extends this discrimination range to 3.5 \\GeVc. Therefore, ${\\bar{p}}/{e^{-}}$ separation can profit from ATC redundancy up to 1-1.5 \\GeVc~and can rely on the ATC up to 3.5 \\GeVc.\\\\ $\\bullet$ {\\bf positrons : }Positrons were also an important issue for AMS-01. They had to be discriminated from a much larger proton flux, with a typical ratio : ${p}/e^{+} \\sim 10^{3}$. Although the ATC design was not optimized for this selection, ${e^{+}/p}$ discrimination could be achieved by using appropriate ATC cuts~\\cite{amsres}, as shown in sec.~\\ref{posi}. \\subsection{Principle of the ATC.} The principle of the ATC counter, used in AMS-01, is based on the \\cher effect to separate $\\bar{p}$ from $e^{-}$ at low energy. Basic relations are recalled here for the reader's convenience. The number of photons created by the \\cher effect, in a material of refractive index $n$, is proportional to : \\begin{equation} N_{\\gamma} \\propto L\\times Z^{2} \\times \\sin^{2} \\theta = L\\times Z^{2} \\times (1-\\frac{1}{n^{2} \\beta^{2}}) \\label{cherequa} \\end{equation} \\noindent where $L$ is the path length in the material, $\\theta$ is the \\cher angle, $Z$ the charge of the incoming particle and $\\beta$ the particle velocity.\\\\ This leads to the following threshold values (in beta or momentum) : \\begin{equation} \\beta_{thres}={1}/n~;~\\mathrm{P}_{thres}=\\frac{mc}{\\sqrt{n^{2}-1}} \\end{equation} \\noindent where $m$ is the mass of the particle at rest.\\\\ The aerogel refractive index ($n=1.035\\pm 0.001$) was chosen \\cite{gougas} to provide a high threshold and a sufficient number of photo-electrons (p.e). The corresponding thresholds, for several particle species, are given in the following table. \\begin{table}[htb] \\begin{center} \\begin{tabular}{|c|c|c|c|c|c|c|} \\hline Particle & $e^{\\pm}$ & $\\pi^{\\pm}$ & p ($\\bar{p}$) & He ($\\bar{He}$) \\\\ \\hline $\\mathrm{P}_{thres}$& 1.91 \\MeVc & 0.52 \\GeVc & 3.51 \\GeVc & 14.0 \\GeVc \\\\ \\hline \\end{tabular} \\end{center} \\end{table} \\hspace{-10mm} \\par \\noindent Electrons and positrons in the 0.5-3.5 \\GeVc~range are far above their threshold, thus giving a full amplitude signal in ATC. In principle, $p$ and $\\bar{p}$ of momentum less than 3.5 \\GeVc, are not expected to give any \\cher signal, thus leading to ${\\bar{p}}/e^{-}$ separation. In the following, this value of 3.5 \\GeVc~will be referred to as the \\cher threshold, i.e. the ATC momentum threshold for antiproton selection. ", "conclusions": "The general behaviour of the AMS apparatus was satisfactory during its first test flight on board the space shuttle Discovery \\cite{amsrep}. The ATC counter allowed ${\\bar{p}}/e^{-}$ separation with a rejection of 330 and an efficiency up to $\\sim 48 \\%$, extending the ${\\bar{p}}/e^{-}$ separation range up to 3.5 \\GeVc. As a secondary result, it has been used, with appropriate cuts, to separate $e^{+}$ from protons with a rejection up to 260 and an efficiency of $\\epsilon_{e^{+}} \\simeq$ 41 \\%. \\newpage \\listoffigures \\newpage" }, "0010/astro-ph0010597_arXiv.txt": { "abstract": "We have performed high-resolution synthesis observations of the $^{12}$CO (J=1-0) line emission from the radio lobe edge-on spiral NGC 3079 using a 7-element mm-wave interferometer at Nobeyama, which consisted of the 45-m telescope and 6-element Array. The molecular nuclear disk (NMD) of 750 pc radius is found to be inclined by 20\\deg\\ from the optical disk, and the NMD has spiral arms. An ultra-high-density molecular core (UHC) was found at the nucleus. The gaseous mass of the UHC within 125 pc radius is as large as $\\sim 3\\times 10^8\\Msun$, an order of magnitude more massive than that in the same area of the Galactic Center, and the mean density is as high as $\\sim 3\\times10^3 {\\rm ~H_2~cm^{-3}}$. A position-velocity diagram along the major axis indicates that the rotation curve starts at a finite velocity exceeding 300 \\kms\\ already from the nucleus. The surface mass density in the central region is estimated to be as high as $\\sim 10^5 \\Msun~{\\rm pc}^{-2}$, producing a very deep gravitational potential. We argue that the very large differential rotation in such a deep potential will keep the UHC gravitationally stable from current star formation. Subject headings: ISM: molecules --- galaxies: individual (NGC 3079) --- galaxies: ISM --- galaxies: kinematics and dynamics --- galaxies: nuclei --- galaxies: structure --- radio lines: ISM ", "introduction": "NGC3079 is an edge-on galaxy known for its pronounced radio, \\ha\\ and X-ray lobes emerging from the nucleus (Hummel et al 1983; Duric et al 1983; Lord et al 1986; Veilleux et al 1994; Pietsch et al 1999). The nuclear outflow may originate with accretion of dense gas onto a compact core (Irwin and Seaquist 1988), and the nucleus exhibits LINER and Seyfert 2 nuclear activity (Lord et al 1986). NGC 3079 has been studied in various radio wavebands in continuum, HI and CO lines as well as maser lines (Irwin and Seaquist 1991; Ford et al. 1986; Duric et al 1983; Duric and Seaquist 1988; Irwin et al 1987; Irwin et al 1988; Young et al. 1988; Sofue and Irwin 1992; Irwin and Sofue 1992; Sawada-Sato et al. 2000). The galaxy's distance is 15.6 Mpc ($H_0=75$ \\kms/Mpc), and the major axis position angle of the optical disk is $167\\Deg$ and the inclination angle 85\\deg. Earlier CO observations with the Nobeyama mm Array (NMA) inferred the existence of a dense nuclear molecular disk, rotating at a high velocity within the central 10$''$ (750 pc) (Sofue and Irwin 1992; Irwin and Sofue 1992). NGC3079 is a galaxy whose unusual nuclear activity can be probed by observing the affected interstellar medium, and can be compared with the Milky Way for its similar edge-on orientation. Knowledge of the kinematics, morphology and density of the interstellar medium in the immediate vicinity of nucleus is particularly important, since the ISM can both fuel the nucleus and trigger activity. The nuclear disk would also play a role in collimating outflows. For this, higher resolution CO observations have been required to determine how the morphology and kinematics of the nuclear molecular disk are related to the nuclear activity. In this paper we report the result of high-resolution CO-line observations of NGC 3079 using a 7-element millimeter wave interferometer at Nobeyama by combining the NMA (Nobeyama mm Array) and 45-m telescope, whose code name was \"RAINBOW\". ", "conclusions": "" }, "0010/astro-ph0010418_arXiv.txt": { "abstract": "We discuss the quantification of the local galaxy population and the impact of the ``New Era of Wide-Field Astronomy'' on this field, and, in particular, systematic errors in the measurement of the Luminosity Function. New results from the 2dFGRS are shown in which some of these selection effects have been removed. We introduce an INT-WFS project which will further reduce the selection biases. We show that there is a correlation between the surface brightness and the luminosity of galaxies and that new technologies are having a big impact on this field. Finally selection criteria from different surveys are modelled and it is shown that some of the major selection effects are surface brightness selection effects. ", "introduction": "Galaxy populations were first studied by Hubble (1926), who developed the familiar Tuning Fork diagram of Ellipticals, Spirals and Barred Spirals. Most bright galaxies can be morphologically classified by their Hubble type. However, many types of galaxy have been found that don't fit the Tuning Fork. These occur both at low redshift and at high redshift where the galaxies can be intrinsically different due to evolution. Some of these galaxies are shown in Fig. 1. The Tuning Fork can be extended to include these new types of galaxy as shown in Fig. 2. \\begin{figure}[h] \\psfig{file=highz1.ps,height=25.0mm,width=25.0mm} \\vspace{-25.0mm} \\hspace{30.0mm} \\psfig{file=highz3.ps,height=25.0mm,width=25.0mm} \\vspace{-25.0mm} \\hspace{5.0mm} \\psfig{file=lowz1.ps,height=25.0mm,width=25.0mm} \\vspace{-25.0mm} \\hspace{5.0mm} \\psfig{file=lsbg2.ps,height=25.0mm,width=25.0mm} \\vspace{50.0mm} \\caption{Some peculiar galaxies. The two on the left hand side are high redshift galaxies in the Hubble Deep Field North. The two on the right are low redshift galaxies from the Millennium Galaxy Catalogue. The farthest right galaxy is a Low Surface Brightness Galaxy.} \\end{figure} \\begin{figure} \\centerline{\\psfig{file=htf.ps,height=55.0mm,width=130.0mm}} \\caption{The various different classes of galaxies that astronomers have named over the years. The classes are positioned roughly in order of magnitude from bright on the left to faint on the right and in surface brightness from high surface brightness at the top to low surface brightness at the bottom. The Hubble Tuning Fork types make up most of the bright galaxies.} \\vspace{-5mm} \\end{figure} However, morphological classification only tells us what type of galaxies there are. It does not tell us what proportion are of each type or whether this varies over time. Galaxy formation and evolution is a relatively new field and so far very little is known about what produced the galaxy population we see today. To understand this process it is first necessary to have some quantitative information about the local galaxy population. The luminosity is an easy to measure quantity which can be used to classify galaxies. The Luminosity Function (LF) (Peebles \\& Hauser 1974) measures the space density of galaxies as a function of luminosity. This can be convolved with different evolutionary models and compared to number - magnitude counts (Driver et al. 1994). However, recent surveys have produced a large range in the measured LF, Fig. 3, with the variation between the surveys much greater than random errors. The systematic errors causing this wide variation must be understood before any progress is going to be made. \\begin{figure} \\centerline{\\psfig{file=lfs2.ps,height=55.0mm,width=130.0mm}} \\caption{This is a plot of eight recent Galaxy Luminosity Functions. The variation at the M$^*$ ($M\\approx-19.5$) point is a factor of 2 and at $M=-14.5$ the variation is a factor of 10. The $M^*$ point is where the statistics are best. The variation would be expected to be of the order 10\\% if it where due to random errors. For more details see Cross et al. (2000).} \\vspace{-5mm} \\end{figure} Disney (1976) pointed out that surface brightness selection effects are important to take into account when measuring the luminosity function. Phillipps \\& Disney (1986) and Phillipps, Davies \\& Disney (1990) went on to calculate the effects of surface brightness. These calculations take into account light lost below the isophote, Malmquist bias and surface brightness dimming due to cosmological expansion. Here, we will discuss how we have used the ``Two Degree Field Galaxy Redshift Survey'' (2dFGRS) to produce a Bivariate Brightness Distribution (BBD), which corrects for these effects. Then we will describe some of the results and some of the selection effects that we have not corrected for. We will then briefly describe the Millennium Galaxy Catalogue (MGC), one of the projects in the Isaac Newton Telescope Wide Field Survey (INT-WFS). The MGC is a deep, imaging survey designed to remove some of the remaining selection effects. The selection effects are then modelled by convolving the BBD with different functions based on the visibility theory of Phillipps, Davies \\& Disney 1990) adopting parameters appropriate for the ESO Slice Project (ESP) (Zucca et al. 1997), the Las Campanas Redshift Survey (LCRS) (Lin et al. 1996) and the Second Southern Sky Redshift Survey (SSRS2) (Marzke et al. 1998). ", "conclusions": "To understand the effects of evolution, it is essential to have a good quantitative knowledge of the local galaxy population. It is apparent from Fig. 4 that our knowledge of the local galaxy population is limited by systematic errors rather than random errors. A major systematic error is caused by surface brightness selection effects. These can be removed using a Bivariate Brightness Distribution. The new technologies in this ``New Era of Wide Field Astronomy'' allow us to detect larger numbers of galaxies with higher precision. These new technologies are the multi-fibre spectrometers such as 2dF and the large format CCDs such as those in the INT Wide Field Camera. These new surveys allow us to properly tackle the selection effects that have dogged our surveys and therefore will allow us to finally pin down the space and luminosity densities of galaxies. \\vspace{-5mm}" }, "0010/astro-ph0010132_arXiv.txt": { "abstract": "This paper is the first one in a series devoted to the study of the dynamics of the atmospheres of long-period variable (LPV) stars. Results from a two-month-long monitoring of the Mira variables RT\\,Cyg and X\\,Oph around maximum light with the ELODIE spectrograph at the Haute-Provence Observatory are presented. The monitoring covers phases 0.80 to 1.16 for RT\\,Cyg and phases 0.83 to 1.04 for X\\,Oph. The cross-correlation profile of the spectrum of RT\\,Cyg with a K0\\,III mask confirms that the absorption lines of RT\\,Cyg in the optical domain appear double around maximum light. No line doubling was found in the optical spectrum of X\\,Oph around maximum light, indicating that this feature is not common to all LPVs. This paper also presents the application to RT\\,Cyg of a new tomographic\\footnote{The word {\\it tomography} is used here in its etymological sense ({\\it `display cuts'}), which differs somewhat from the broader sense in use within the astronomical community (reconstruction of a structure using projections taken under different angles).} technique deriving the velocity field across the atmosphere by cross-correlating the optical spectrum with numerical masks constructed from synthetic spectra and probing layers of increasing depths. This technique reveals that both the temporal evolution of the line doubling, and its variation with depth in the atmosphere of RT\\,Cyg, are consistent with the `\\sch\\ scenario'. This scenario relates the temporal evolution of the red and blue peaks of the double absorption lines to the progression of a shock wave in the atmosphere. The temporal evolution of the Balmer \\ha, \\hb, \\hg\\ and \\hd\\ emission lines around maximum light is also presented for RT\\,Cyg and X\\,Oph. The velocity variations of \\ha\\ and of the absorption lines are discussed in the framework of two competing models for the formation of Balmer emission lines in LPV stars. ", "introduction": "Long-period variable stars (LPVs) are cool giant stars showing more or less periodic light variations with amplitudes of several magnitudes in the visual and with periods of several hundred days. Depending on their visual amplitudes and on the regularity of their variability cycles, they appear in several flavours, namely Mira Ceti-type variables (Mira stars or Miras), Semi-Regular variables (of the SRa or SRb subtypes) or Irregular variables (of the Lb subtype). Mira and SR variables represent one of the latest stages in the evolution of stars with initial masses in the approximate range 1 to 9~\\Msun. It is known since long that the brightness variations of LPVs go along with spectral changes: (i) hydrogen and some metallic lines turn from absorption to emission after minimum light (Merrill 1921), (ii) the velocity of the emission and absorption lines correlates with phase and excitation potential (Merrill 1923a; Adams 1941), (iii) several absorption lines appear double around maximum light (Adams 1941; Merrill \\& Greenstein 1958; Maehara 1968). In a review on red stars, Merrill (1955) was the first to suggest that the bright emission lines sometimes appearing in LPVs may be explained by some kind of ``hot front'' moving outward. He further suggested that this running hot front may have the structure of a shock wave. Although the shock wave scenario is nowadays quite widely accepted (see de la Reza 1986 and references therein), the lack of self-consistent pulsation models for LPV stars (due to the important role played by convection and the difficulty of modelling it) prevents from grasping exactly how, why and where emission and double absorption lines form. As a consequence, several (sometimes conflicting) theories aiming at explaining the spectral peculiarities of LPVs have appeared in the literature. Although the most popular models locate the formation of emission lines in the hot wake of the shock (e.g., Gillet 1988a), an alternative model proposed by Magnan \\& de Laverny (1997) associates emission lines with purely radiative non-LTE processes independently of any shock wave. In this model, the double-peak nature of the emission lines is not connected in any way with the velocity field, contrary to the shock-wave model. Concerning the absorption lines, their doubling around maximum light is most easily seen in the rather clean near-infrared spectral domain (e.g., Gillet et al.\\ 1985; Hinkle et al.\\ 1997 and references therein). Long-term radial velocity measurements of the infrared rotation-vibration lines of CO have revealed well-defined variations with phase, regularly repeating from one cycle to the next and following a typical S-shaped curve (Hinkle et al.\\ 1997). These radial-velocity variations are generally believed to reflect the differential bulk motions occurring in the large and tenuous atmosphere of LPVs and associated with its pulsation. Alternatively, the complex absorption profiles have also been explained without resorting to differential velocity fields: Gillet et al.\\ (1985) have argued that the doubling of the metallic absorption lines results in fact from the development of an emission core in the line. This emission component is caused by the radiative release of the thermal energy dumped into the post-shock layer by the shock front. This model has the advantage of being able to account for the P~Cygni line profiles sometimes observed in LPV spectra (and appearing when the emission core becomes brighter than the local continuum). A similar explanation in terms of the shock-induced temperature inversion has been proposed by Karp (1975) for the line doubling observed in Cepheid variables. Detailed radiative-transfer calculations have shown that the formation of double absorption lines does not necessarily require a velocity gradient, since a temperature inversion is in principle sufficient. Recent studies have clearly validated theories accounting for the complex absorption line profiles observed in several kinds of pulsating stars (RR\\,Lyrae: Chadid \\& Gillet 1996, Fokin \\& Gillet 1997; $\\beta$\\,Cephei stars: Mathias et al.\\ 1998; BL\\,Herculis: Gillet et al.\\ 1994, Fokin \\& Gillet 1994; RV\\,Tauri: Gillet et al.\\ 1989a) in terms of velocity fields. The validity of the \\sch\\ scenario accounting for double absorption lines in terms of a velocity gradient in the photosphere remains however to be demonstrated for LPV stars. The present paper presents the first clear observational evidence thereof. ", "conclusions": "The origin of double-peaked emission and absorption lines in the complex LPV atmospheres is a long-standing problem. If the shock wave and the associated velocity gradient are generally believed to be at the origin of the complex line profiles of LPV stars (de la Reza 1986; Gillet 1988b; Hinkle et al.\\ 1997), alternative models not resorting to differential atmospheric motions are sometimes invoked. They involve either purely non-LTE effects to account for the double-peaked Balmer emission (Magnan \\& de Laverny 1997), or a temperature inversion to account for the splitting of the absorption lines (Karp 1975; Gillet et al.\\ 1985). Thanks to the cross-correlation technique and a daily survey of the Mira star RT\\,Cyg, the role of velocity fields in the formation of genuinely double absorption lines has been undoubtedly confirmed in the framework of the \\sch\\ mechanism. Evidence for the existence of strong shock waves in the outer layers of LPVs becomes compelling, although not all Mira stars exhibit the \\sch\\ scenario, as demonstrated by the observations of X\\,Oph collected in the present paper. The origin of this distinct behaviour will be investigated in the next paper of this series. A tomographic technique that opens new perspectives for the study of the dynamics of LPVs has been used to visualize the outward motion of the shock wave in the atmosphere of RT\\,Cyg. The evolution of the \\ha, \\hb, \\hg\\ and \\hd\\ emission-line profiles during 35\\% (20\\%) of the cycle of RT\\,Cyg (X\\,Oph) around maximum light has also been discussed. The comparison of this data set with the temporal evolution of the profile of absorption lines offers interesting perspectives to identify the origin of the complex profile of the \\ha\\ emission line. The present paper had no other ambition than providing a description of the data collected with the ELODIE spectrograph around maximum light for RT\\,Cyg and X\\,Oph. These data may be used as constraints for future detailed models. Firm conclusions on the line-formation mechanisms in Mira variables should thus await the advent of such detailed dynamic models coupling hydrodynamics to radiative transfer." }, "0010/astro-ph0010654_arXiv.txt": { "abstract": "In a previous paper on the Magellanic Clouds we demonstrated that coeval clusters provide a powerful tool for probing the progenitor masses of Wolf-Rayet stars (W-Rs) and Luminous Blue Variables (LBVs). Here we extend this work to the higher metallicity regions of the Milky Way, studying 12 Galactic clusters. We present new spectral types for the unevolved stars and use these, plus data from the literature, to construct H-R diagrams. We find that all but two of the clusters are highly coeval, with the highest mass stars having formed over a period of less than 1~Myr. The turn-off masses show that at Milky Way metallicities some W-Rs (of early WN type) come from stars with masses as low as 20--25$\\cal M_\\odot$. Other early-type WNs appears to have evolved from high masses, suggesting a large range of masses evolve through an early-WN stage. On the other hand, WN7 stars are found only in clusters with very high turn-off masses, $>120\\cal M_\\odot$. Similarly the LBVs are only found in clusters with the highest turn-off masses, as we found in the Magellanic Clouds, providing very strong evidence that LBVs are a normal stage in the evolution of the most massive stars. Although clusters containing WN7s and LBVs can be as young as 1~Myr, we argue that these objects are evolved, and that the young age simply reflects the very high masses that characterize the progenitors of such stars. In particular we show that the LBV $\\eta$~Car appears to be coeval with the rest of the Trumpler~14/16 complex. Although the WCs in the Magellanic Clouds were found in clusers with turn-off masses as low as $45\\cal M_\\odot$, the three Galactic WCs in our sample are all found in clusters with high turn-off masses ($>70\\cal M_\\odot$); whether this difference is significant or due to small-number statistics remains to be seen. The BCs of Galactic W-Rs are hard to establish using the cluster turn-off method, but are consistent with the ``standard model\" of Hillier. ", "introduction": "\\subsection{Background} Massive, luminous stars begin their H-burning lives as hot, O-type stars. During their main-sequence evolution (2.5--8~Myr for stars with initial masses of 120--20$\\cal M_\\odot$) they may lose a significant amount of their mass due to strong stellar winds. The observed mass-loss rates suggest that the highest-mass stars will lose as much as half their mass during the H-burning stage. Since these winds are driven by radiation pressure through highly-ionized metal lines, the mass-loss rates will increase with stellar luminosity, metallicity, and effective temperature. It was Conti (1976) who first suggested that mass-loss provided a simple explanation for how Wolf-Rayet (W-R) stars form. In the modern version of the ``Conti scenario\" (Maeder \\& Conti 1994), this mass loss results in a stripping off of the H-rich outer layers of the star, resulting in a WN-type W-R star, in which the H-burning products (He and N) are enriched at the surface, with strong, broad emission lines indicative of enhanced stellar winds forming an extended atmosphere. Most WNs are presumed to be He-burning objects, although there is evidence that a few H-rich late-type WNs are still near the end of the core-H burning phase (Conti et al.\\ 1995). Further mass-loss and evolution reveals the products of He-burning (C and O) at the surface, and the star is spectroscopically identified as a WC-type W-R. Possibly the highest mass stars also go through a Luminous Blue Variable (LBV) stage on their way to becoming W-Rs, with the large, episodic mass loss that characterizes LBVs aiding the process. Stars of slightly lower luminosity (mass) will not have an LBV phase, but recent observations (Massey 1998a) suggest that they do go through an intermediate red supergiant (RSG) phase, even at high metallicities, albeit for a short time. At some lower luminosity one expects that the mass-loss rates are sufficient to produce a WN-type W-R, but not a WC. And, at even lower luminosities, mass-loss rates are so low that the He-burning stage is spent entirely as a red supergiant (RSG) and not as a W-R star. In this version of the ``Conti scenario\" we thus might expect the following ``paths\" to be followed, in order of decreasing luminosities: \\begin{centering} O$\\longrightarrow$LBV$\\longrightarrow$WN$\\longrightarrow$WC (1) O$\\longrightarrow$RSG$\\longrightarrow$WN$\\longrightarrow$WC (2) O$\\longrightarrow$RSG$\\longrightarrow$WN (3) OB$\\longrightarrow$RSG (4) \\end{centering} The masses corresponding to these paths are unknown (and indeed we are unsure even of the qualitative correctness of these paths), but ``standard guesses\" for characteristic values would be 1: $\\geq80\\cal M_\\odot$, 2: $60\\cal M_\\odot$, 3: $40\\cal M_\\odot$, and 4: $20\\cal M_\\odot$. We emphasize that these are purely speculative values, and that the actual ranges should depend upon metallicity. Indeed it is to address this issue that the present series of papers has come about. Between the initial O-type stage (of luminosity class ``V\") and the He-burning stage (LBVs, W-Rs, RSGs) the star will become a B-type supergiant; most, but not all, of these are expected to still be H-burning objects (Massey et al.\\ 1995b). The luminous A-F supergiants are very short-lived intermediate stages during He-burning for stars of intermediate high-mass, depending upon the metallicity. The precursor to SN1987A is believed to have been a ``second-generation\" B-type supergiant, a He-burning object of somewhat lower mass than those being discussed here. In addition to the question of whether or not the above paths are correct, and what masses to assign to each as a function of metallicity, we are also interested in the evolutionary significance, if any, to the Wolf-Rayet spectral subtypes. WN-type W-R stars are classified as ``early\" (WNE) or ``late\" (WNL) depending on whether NV~$\\lambda\\lambda 4602,19$ or NIII~$\\lambda\\lambda 4634,42$ dominates; e.g., WNEs correspond to numerical subtypes WN2, WN3, WN4, while WNLs consist of subtypes WN6, WN7, WN8, and WN9, with the WN5 stars split between the two groups. Similarly WC-type W-R stars are classified as WCE or WCL depending upon whether CIII~$\\lambda 5696$ dominates over CIV~$\\lambda 5808$; e.g., corresponding to spectral subtypes WC4 through WC6 and WC8-9, with some WC7s falling into each camp (Conti \\& Massey 1989). Do these subtypes mean anything in an evolutionary sense? Various authors have claimed so (see, for example, Moffat 1982), but this conjecture does not seem borne out by either observational or theoretical studies. Our understanding of massive star evolution is limited, in part, because of the difficulty of constructing stellar models from first principles. The physics of massive stars is complicated by strong stellar winds, and the choice of the functional dependence of mass-loss rates on stellar parameters (luminosity, temperature, mass, and metallicity) greatly influences the theoretical tracks (e.g., Meynet et al.\\ 1994), particularly in the later stages of evolution. In addition, the models are sensitive to the amount of mixing. However, there is little agreement on the treatment of the relevant processes of semi-convection and overshooting (Maeder \\& Conti 1994), while the most recent work has emphasized the significant role that rotation may play in this regard (Maeder 1997, 1999). Nevertheless, empirical studies of massive star evolution provide confidence that the above picture is correct, and are beginning to provide quantitative information on the mass ranges corresponding the various paths. These studies provide an observational basis against which the models can be evaluated and refined. (For a humorous rendering of this process, the reader is referred to Fig.~5 in Conti 1982.) \\subsection{An Observational Approach} \\subsubsection{Global Studies} The galaxies of the Local Group provide perfect laboratories for pursing these studies observationally, as the metallicity differs by almost an order of magnitude (SMC to M31) among the galaxies currently active in forming stars. During the past few years there have been a number of studies of {\\it mixed-age} populations, the relative number of {\\it this} and {\\it that}. The implicit assumption of these studies are that the IMF slopes are identical and cover regions that provide a good sampling of stellar stages over time. The number ratios provide quantitative criterion for the models to attempt to match. These studies have found the following: (1) The relative number of RSGs to W-Rs decreases with increasing metallicity (Massey 1998a; Massey \\& Johnson 1998). Histograms of the number of RSGs vs.\\ luminosity reveal that there are proportionately fewer high luminosity RSGs at higher metallicity, while the lack of a sharp luminosity cut-off supports the interpretation that as $Z$ increases, massive stars spend a greater fraction of their He-burning phase as W-Rs, and a smaller fraction as RSGs, rather than there being a difference in the actual mass ranges that go through a RSG phase. This is why we indicated a RSG phase at high luminosities (path 2 above). Possibly even LBVs will go---or have gone---through a RSG phase, but this is unknown. We also do not know if massive stars go through a RSG phase at the highest metallicity: the relation between the RSG/W-R ratio and metallicity appears to flatten below the high metallicity that characterize M31, but only a few regions have been surveyed in that large galaxy, and more data are being gathered to resolve this. (2) The relative number of WCs to WNs increases with increasing metallicity, with the notable exception of the star-burst galaxy IC~10 (Fig.~8 in Massey \\& Johnson 1998). This trend is also in accord with the predictions of the Conti scenario, as increased mass-loss makes it possible for a star of a given luminosity to reach the WC stage sooner, spending more of its W-R stage as a WC than a WN. (The explanation for IC10's peculiarly high WC/WN ratio remains a mystery at present. See discussion in Massey \\& Johnson 1998.) (3) Significant differences in the spectral subtypes found in the Magellanic Clouds compared to the Milky Way were noted by Smith (1968): no WCLs are found in the LMC or SMC, and most of the ones known in the Milky Way are found inwards of the solar circle. Similar differences are seen for the WNs. Armandroff \\& Massey (1991) showed that the WC line widths (which are correlated with spectral subclass) change systematically with metallicity, extending an important finding by Willis, Schild, \\& Smith (1992) to other galaxies of the Local Group. Massey (1996) proposed that the WC spectral subtypes are nothing more than an atmospheric effect due to metallicity. (See also Massey \\& Johnson 1998). Recently Crowther (2000) has demonstrated from W-R model atmospheres that the WN subtypes may similarly be a reflection of metallicity rather than other stellar parameters, at least in terms of WN3 through WN6. \\subsubsection{Coeval Associations} A more direct way exists to attack the problem of understanding massive star evolution. By using coeval associations that contain evolved, massive stars, we can in fact directly {\\it measure} the mass ranges that correspond to the above evolutionary paths as a function of metallicity. This is the subject of the current series of papers. In Paper I of this series (Massey, Waterhouse, \\& DeGioia-Eastwood 2000) we established that many Magellanic Cloud OB associations and young clusters are sufficiently coeval ($\\Delta \\tau<$1~Myr) that we can measure a meaningful cluster turn-off mass; e.g., the mass of the most massive star on the main-sequence. These turn-off masses then place a lower-limit on the mass of the progenitors of the evolved stars in the cluster, to the extent that star-formation proceeded coevally. If the cluster is well populated, than the initial mass of the turn-off star is a good approximation to the initial mass of the progenitor of the W-R star. In addition, the bolometric luminosity of the cluster turn-off sets useful limits on the bolometric corrections for the evolved stars, allowing tests of Wolf-Rayet model atmospheres, such as Hillier's ``standard model\" (Hillier 1987, 1990). The results were quite revealing. In a study of 19 OB associations in the Magellanic Clouds, we found that at the low metallicities that characterize the SMC, only the highest-mass stars ($>70 \\cal M_\\odot$) become Wolf-Rayet stars, although the sample is small. This is equivalent to saying that path (3) above occurs only for $M>70\\cal M_\\odot$ for $Z\\leq 0.005$.\\footnote{For convenience in talking about the metallicity $Z$, we adopt $Z=0.018$ for the solar neighborhood, corresponding to log~O/H+12=8.70 (Estebam \\& Peimbert 1995). If we then simply scale $Z$ relative to the easily-measured oxygen abundance, $Z=0.005$ for the SMC (log~O/H+12=8.13) and $Z=0.008$ for the LMC (log~O/H=8.37), e.g., Russell \\& Dopita (1990). Although it is well recognized that different metals will have different relative abundances, it is fortuitous that it is oxygen (along with carbon and nitrogen) which are the primary accelerators of the stellar winds at the high effective temperatures appropriate to O-type stars (Abbott 1982).} At the higher metallicity of the LMC ($Z=0.008$), WN W-R stars come from stars with masses as low as $30\\cal M_\\odot$. We also found that WC stars are found in the same clusters as WNEs; e.g., the lowest turn-off mass found for a cluster containing WC stars was 45$\\cal M_\\odot$, suggesting that stars with masses from 30--45$\\cal M_\\odot$ might correspond to path (3), while stars with masses 45-85$\\cal M_\\odot$ correspond to path (2). The rare ``Ofpe/WN9\" stars\\footnote{The stars may be rarer at higher metallicity than at low; ten are known in the LMC (Bohannan \\& Walborn 1989), while only one is known in the Milky Way (Bohannan \\& Crowther 1999). Six are known in M~33 (Massey et al.\\ 1996), while one is known in the higher metallicity M~31 (Massey 1998b).}, once thought to be a transition type between ``Of\" stars and Wolf-Rayet stars, are only found in clusters and associations with the lowest turn-off masses, 25-35$\\cal M_\\odot$. Recently the Ofpe/WN9 stars were implicated in the LBV phase of massive stars, after one Ofpe/WN9 star (R127) underwent an ``LBV-like\" outburst. But, the classical LBVs in our LMC sample, including the archetype itself S Doradus, are found in clusters with the very highest turn-off masses, $>85 \\cal M_\\odot$. (Similarly the SMC W-R star HD~5980, which many consider to be a ``true\" LBV [Barb\\'a et al.\\ 1995] is found in a cluster with a very high turn-off mass.) We conclude that the Ofpe/WN9 stars are just the lowest-mass versions of Wolf-Rayet stars. True LBVs, on the other hand, are found only in the clusters with the highest mass turn-offs, suggesting that they are indeed stars near their Eddington limit and are a normal stage of the most massive stars. Our study also shed light on the origin of the different W-R classes and subtypes, at least at the modest metallicity that characterizes the LMC. The early-type WN stars (WNEs) in the LMC are found in clusters with a large range of turn-off masses (from 30$\\cal M_\\odot$ to 100$\\cal M_\\odot$ or more), suggesting that these are a stage that most massive stars go through at LMC-like metallicities. We turn now to the higher metallicity of our own Milky Way, and pose the question of where do LBVs and Wolf-Rayets of various types come from at a metallicity considerably higher than that of the Magellanic Clouds. ", "conclusions": " \\begin{enumerate} \\item Wolf-Rayet stars in the Milky Way are found in clusters containing a large range of turn-off masses. The data suggest that at the metallicity that characterizes the Milky Way some early-type WN Wolf-Rayet stars come from progenitors with masses as low as 20--25$\\cal M_\\odot$. \\item The WNEs may come from a large range in masses, as they do in the Magellanic Clouds (Paper~I), but this result is uncertain, as the one high-mass WNE star in our sample is a member of a close binary, V444~Cyg. \\item WN7 stars are found only in clusters with the highest masses. The youngest of these are only 1~Myr old. Although these could still be H-burning objects, ``guilt by association\" suggests that these are in fact evolved massive stars, as the youngest region also contains the LBV $\\eta$~Car, thought to be an evolved object. \\item $\\eta$ Car itself is found to be highly coeval with the rest of the Tr~14/16 complex, despite the region's young age. It, and the LBV candidate VI~Cyg~No.~12, are found in clusters with the highest masses. This is identical to what we found for the Magellanic Clouds in Paper~I, and argues strongly that LBVs are a normal stage in the evolution of the most massive stars. \\item The Galactic WC stars are found in clusters with turn-off masses $> 70 \\cal M_\\odot$. In the Magellanic Clouds we find WCs occurring in clusters with masses as low as 45~$\\cal M_\\odot$. We argued in Paper~I that most WNs thus evolve to WCs. The data for the Milky Way might suggest that only the more massive stars become WCs, but the sample size in the Milky Way is small (3 WCs in 3 clusters) and additional data are needed. \\item The BCs of Galactic W-Rs are hard to determine using the cluster turn-off methods, as considerable luminosity evolution is expected at the higher mass-loss rates that characterize the Milky Way luminous stars. The data are at least consistent with the ``standard model\" of Hillier (1987, 1990) as applied to two of the stars in our sample (Crowther et al.\\ 1995). In Paper~I we found concluded that there was excellent agreement, with the BCs of early WN and WC stars found to be extreme ($\\approx -6$~mag). \\end{enumerate} We note that much recent work has established the need to extend the theoretical evolutionary tracks to masses higher than 120$\\cal M_\\odot$. Stars with masses estimated to be as high as 160$\\cal M_\\odot$ have been found in the R~136 cluster (Massey \\& Hunter 1998), and the Galactic clusters Trumpler~14/16, C1715-387, Pismis~24, and Cyg~OB2 all contain stars whose luminosities place them above the highest evolutionary track computed by the Geneva group (120~$\\cal M_\\odot$). Future observational work is needed to extend the H-R diagrams of these and other Galactic clusters, and to investigate other coeval regions in the Milky Way and nearby galaxies that can be used to extend these studies." }, "0010/astro-ph0010462_arXiv.txt": { "abstract": "We present preliminary results on the effects of mergers on the chemical properties of galactic objects in hierarchical clustering scenarios. We adopt a hydrodynamical chemical code that allows to describe the coupled evolution of dark matter and baryons within a cosmological context. We found that disk-like and spheroid-like objects have distinctive metallicity patterns that may be the result of different evolution. ", "introduction": " ", "conclusions": "" }, "0010/astro-ph0010181_arXiv.txt": { "abstract": "We have imaged seventeen recently discovered Sirius-like binary systems with {\\it HST/WFPC2} and resolved the white dwarf secondary in eight cases. Most of the implied orbital periods are of order several hundred years, but in three cases (56 Per, $\\zeta$ Cygni and RE~J1925$-$566) the periods are short enough that it may be possible to detect orbital motion within a few years. It will then be possible to derive dynamically determined masses for the white dwarfs, and potentially these stars could be used as stringent tests of the mass-radius relation and initial-final mass relation. ", "introduction": "Although universally accepted, the white dwarf mass-radius relation (Chandrasekhar 1931; Hamada \\& Salpeter 1961; Wood 1992) has resisted in-depth examination through direct observations for $\\approx$70 years. The difficulty arises mainly from obtaining accurate, model-independent measurements of the masses and radii of known white dwarfs. A combination of spectroscopically derived temperatures and gravities, and accurate parallaxes can provide the independent means of determining $M_{\\rm WD}$ and $R_{\\rm WD}$, and therefore a new route for examining evolutionary models. The white dwarfs observed by {\\it Hipparcos} (Vauclair et~al. 1997; Provencal et~al. 1998) were entirely consistent with expectations, although they were restricted mainly to a narrow range around $0.6M_\\odot$. Furthermore, even with these new parallaxes the overall uncertainties are still too large to provide a really thorough examination of the differences between theoretical models which assume a variety of core and envelope compositions. White dwarfs in resolved binary systems can provide the most stringent tests of evolutionary models, since the mass can be determined from the orbital and physical elements of the system. In practise, though, few such systems are available to be studied in sufficient detail (they are: Sirius, Procyon, 40 Eri and Stein 2051). However, as reported at previous workshops (e.g.~Burleigh 1999) $>20$ new Sirius-like binaries have been identified through the {\\it ROSAT WFC} and {\\it EUVE} all-sky surveys. Each system consists of a normal or subgiant star plus a hot white dwarf companion, which is responsible for the soft X-ray and EUV flux detected by {\\it ROSAT} and {\\it EUVE}. Crucially, in most cases, the bright primary has an accurate {\\it Hipparcos} parallax, yielding a precise distance for the white dwarf. Thus this new sample of Sirius-like binaries presents a golden opportunity of extending the sample of well-studied white dwarf binary systems. Our first aim, therefore, is to identify those systems that can be resolved and have their orbits measured within an acceptable timescale. Unfortunately, none of these new binaries can be resolved from the ground, due to the huge difference in brightness between the components ($>5$ mags.), although we know from radial velocity studies that most of these systems are wide with periods $>$few years (e.g.~Vennes, Christian, \\& Thorstensen 1998). The {\\it Hubble Space Telescope} provides an answer to this problem. Unencombered by the Earth's atmosphere, it allows imaging in the UV where the brightness of the hot white dwarf and its companion are similar. It also delivers a diffraction limited resolution of $\\approx0.05\\arcsec$, making it possible to resolve binary components with separations as small as 0.1\\arcsec. We report here the first results of a snapshot survey with {\\it HST/WFPC2} to image these new Sirius-like binary systems. ", "conclusions": "" }, "0010/astro-ph0010148_arXiv.txt": { "abstract": "We present large-scale CCD images of a dozen low surface brightness objects in the vicinity of giant Sc galaxy NGC~6946, which have been found recently on POSS-II copies. Six of them: UGC~11583, kk251, kk252, kkr55, kkr56, and kkr59 are dwarf irregular galaxies resolved into stars. We determined their distances from the luminosity of the brightest blue stars. The mean estimated distance to the NGC~6946 group is 5.9$\\pm$0.4 Mpc. Together with Cepheus 1, discovered before, the group consist of 8 late type galaxies. Almost all of them are detected or marginally detected in the HI line. The group is characterized by a linear projected diameter of 420 kpc, a radial velocity dispersion of 95 km~s$^{-1}$, and a virial mass-to-luminosity ratio about 56 $M_{\\sun}/L_{\\sun}$. ", "introduction": "NGC 6946 is amongst the seven more massive galaxies in the Local Volume (=LV) with distances $D <$ 7 Mpc (Karachentsev et al. 1999). 51\\% of the population of the LV are within the gravitational influence of these \"oligarchs\". Each of 7 the giant galaxies, except NGC~6946, has a lot of dwarf companions: from 7--9 members around Sc galaxies M101 and NGC~5236 to 32 members around giant elliptical galaxy NGC~5128. In this sense the absence of accompying dwarf systems around NGC~6946 within a radius of $\\sim$1 Mpc (Tully, 1988) looked like a remarkable anomaly. However, the apparent isolation of NGC~6946 may be related with its location on a border of the Local Void (Tully 1988). The substantial galactic extinction in the direction of NGC~6946 has also an influence on the visibility of faint dwarf galaxies. Based on copies of the Second Palomar Observatory Sky Survey (=POSS-II), Karachentseva \\& Karachentsev (1998) undertook a search for galaxies of low surface brightness in the vicinity of normal galaxies in the LV. Observations of these galaxies (kk- list) with the 100-m Effelsberg radio telescope revealed three dwarf galaxies: UGC~11583 = kk250, kk251, and kk252, with positions and radial velocities close to NGC~6946 (Huchtmeier et al. 1997). Using the luminosity of the brightest stars in NGC~6946 and UGC~11583, Sharina et al. (1997) derived distance moduli to these galaxies: 28.90 and 29.13 mag, respectively. During the course of a search for compact high velocity clouds Burton et al. (1999) discovered beside NGC~6946 a low surface brightness galaxy with radial velocity $V_h= +58$ km~s$^{-1}$ named by them Cepheus 1. Subsequent searches for nearby dwarf galaxies by Karachentseva et al. (1999) (kkr- list) and Karachentseva et al. (2000) (kkh- list) led to the discovery of some more LSB objects around NGC~6946. In this paper we present results of optical and HI observations of these new objects, particularly data on three new probable companions of NGC~6946. ", "conclusions": "Being located at a low galactic latitude, $b = 12\\degr$, in a zone of substantial extinction, NGC~6946 looks at first glance like an isolated giant spiral galaxy on the border of the Local Void (Tully, 1988). Recent searches for low surface brightness galaxies on the POSS-II films revealed about a dozen probable companions to NGC~6946. Subsequent HI observations of the candidates and distance estimates from the brightest stars showed that at least 7 dwarf irregular galaxies can be considered as physical companions of NGC~6946. Therefore, a new nearby group of galaxies has been discovered at a distance of (5.9$\\pm$0.4) Mpc. The mean projected diameter of the group is 420 kpc, its root mean square radial velocity is 95 km~s$^{-1}$, and the virial mass-to-luminosity ratio is $\\sim56 M_{\\sun}/L_{\\sun}$, typical for small galaxy groups. The two nearby giant face-on Sc galaxies: M101 and NGC~6946, and their dwarfs are alike with respect to their structure and kinematics. Being gas-rich galaxies, the members of the NGC 6946 group need to be studied in more detail with radio HI synthesis interferometry." }, "0010/astro-ph0010524_arXiv.txt": { "abstract": "We present a detailed analysis of the iron K$\\alpha$ line variabilities in NGC 4151 by using long ASCA observation data obtained in May 1995. Despite the relatively small amplitude variations in the continuum flux, the iron K$\\alpha$ line flux and profile show dramatic variations. Particularly, the line profile changes from single peak to seeming double peaks and back in time scales of a few 10$^4$ sec. The seemingly double-peaked profiles can be well interpreted as line emission from a Keplerian ring around a massive black hole. An absorption line at around 5.9 keV is also marginnaly detected. We discussed current Fe K line models, but none of them can well explain the observed line and continuum variations. ", "introduction": "Broad Fe K fluorescence line profiles, detected in many AGNs (Tanaka et al. 1995, Nandra et al. 1997, Wang et al. 1999a, Nandra et al. 1999, and references therein), are thought to arise from the accretion disk near the center massive black hole of AGNs. The line profile, thus, carries important information about the distribution and kinematics of optically material as well as about the space properties in the vicinity of the postulated super massive black hole. Doppler and gravitational shifts would imprint characteristic signatures on the line profile which map the geometric and dynamical distributions of matter surround the black hole. Additional information concerning the geometry of X-ray source and matter in the active nucleus could in principle be derived by studying the rapid variability (on time scales of several 10$^4$ s or less) of the line profile, intensity and their relationship with the continuum variations. So far, rapid variability in Iron K line has been detected in NGC 7314 (Yaqoob et al. 1996), MCG --6-30-15 (Iwasawa et al. 1996, 1999), NGC 4051 (Wang et al. 1999b) and NGC 3516 (Nandra et al. 1999). The observed Fe K$\\alpha$ line rapid variabilities in different targets or during different observations of the same target are quite different. Yaqoob el al. (1996) presented the evidence for rapid variability of the Fe K line profile in the narrow-line Seyfert galaxy NGC 7314, which is consistent with a disk-line of constant equivalent width superposed on a constant flux narrow line (presumably from the torus). Iwasawa et al. (1996) discovered that, during the ASCA observation on MCG --6-30-15 in 1994, when the source is bright, the Fe K line is weak and dominated by the narrow core, whilst during a deep minimum, a huge red tail appears. The intensity of broad Fe K line correlates inversely with the continuum flux on the time scales of several 10$^4$ s. And Iwasawa et al. (1999) discovered that during the ASCA observation on the same target in 1997, the Fe K$\\alpha$ line of a flare phase peaks around 5keV and most of its emission is shifted to below 6 keV with no component detected at 6.4keV. While for NGC 4051, during the ASCA observation in 1994, the equivalent width and the width of Fe K line correlate positively with the continuum flux (Wang et al. 1999b), which shows an opposite trend with the ASCA observation on MCG --6-30-15 in 1994. The rapid variabilities of Fe K$\\alpha$ line profiles in NGC 3516 are also quite irregular (Nandra et al. 1999, Wang et al. 2000). The nearby Seyfert 1.5 galaxy (z = 0.0033), NGC 4151, was first established as a bright X-ray source some thirty years ago (Gursky et al. 1971). Since that time, this archetypal low-luminosity active galaxy has invariably been considered as a prime target by all major X-ray astronomy missions. In 1995, NGC 4151 was observed by ASCA from May 10 to 12 (Leighly et al. 1997). The net exposure time of the observation is about 200 ks. Because of the enhanced sensitivity provided by ASCA and the much longer exposure time, the obtained Fe K$\\alpha$ line profile has very high data quality (Wang et al. 1999a, here after W99). In this letter we report the discovery of unique and rapid variabilities of the iron line profiles in NGC 4151 during the ASCA observation in 1995. ", "conclusions": "\\subsection{The seemingly double-peaked line profiles} Massive black holes are generally thought to exist at the centers of active galaxies (Rees 1984), but unambiguous identification of a black hole has been impeded by lack of evidence for the strong-field relativistic effects expected in the vicinity of black holes. Because of its unique anticipated profiles, the Fe K$\\alpha$ fluorescence line has long been considered as an observational key to the study of the innermost regions of active galactic nuclei (AGN). Fabian et al. (1989) and Laor (1991) pointed out that the line profiles from the accretion disk should be broad, skewed and double-peaked (dependent on the inclination of the disk and the line-producing region in the disk), which are characteristics of the Doppler effects of the accretion disk and the strong gravitational field in the vicinity of the central black hole. The Fe K$\\alpha$ line profiles detected in many AGNs have been proved to be skewed and extremely broad (Tanaka et al. 1995, Nandra et al. 1997, Wang et al. 1999a, Nandra et al. 1999, and references therein), consistent with the disk line model, so give the ever strongest evidence for the presence of massive black holes and accretion disks in the center of AGNs. However, some alternative models, such as Comptonization and jet, can also explain the broad and skewed Fe K$\\alpha$ line profiles. Those models do not require strong gravitational field and so give no support to the presence of massive black holes (Fabian et al. 1995, Misra \\& Sutaria 1999). But different to the disk line model, those models predict a smooth line feature, but not a double-peaked one. Though some indirect evidences have shown that they are not satisfactory (Fabian et al. 1995, Misra 2000, Reynolds \\& Wilms 2000), as Misra \\& Kembhavi (1998) have pointed out, we infer that the detection of a double-peaked line profile is the key to confirm the disk line model for the observed Fe K$\\alpha$ lines, and so on confirm the detection of extremely strong gravitational field. In this paper we find that the Fe K$\\alpha$ line profile of NGC 4151 is seemingly double-peaked during phase I-3 and I-4 in Fig1. Because an additional resonant absorption line at 5.9 keV has also been detected in the line profile in I-4, we only focus on the profile of I-3 in this section. We show the seemingly double-peaked line profile of I-3 in detail in Fig.3, and the best-fit two gaussians and disk-line models are also plotted. The seemingly double-peaked line profile strongly suggest that the line arises from materials with disk-like geometry. Additional strong gravitational field is also needed to explain the red shift of the profile. Others models for the broad line profiles, such as Comptonization or jet, can not be responsible for a double-peaked line profile, and disk-like outflows or inflows with high velocity can not be responsible for the red shift of the line profile. The best disk line fits show that the line mainly arises from a narrow ring between 10.8 R$_g$ and 18.6 R$_g$. Currently, perhaps due to the limited energy resolution and data quality, obvious double-peaked Fe K$\\alpha$ line profile has never been detected in any active galactic nuclei. We noticed that the line profile of MCG --6-30-15 in Fig. 1b of Wang et el. (1999a) looks double-peaked, but the plot was folded through the effective area of the detectors which rises sharply towards the low energy, so did not give the true profile of the line. While in plottings corrected for detectors' effective area, we can not see obvious double peaks (see Fig. 7 of Iwasawa et al. 1996 and Fig. 6 of Fabian et al. 2000). We hope that the new generation X-ray satellites, such as XMM, can show us clearly double-peaked iron line profiles of AGNs. \\subsection{The rapid variabilities of the Fe K$\\alpha$ line profiles} Additional to the double-peaked line profiles, the rapid variabilities of the Fe K line profiles in NGC 4151 are also detected in this letter. We divided the whole ASCA observation of NGC 4151 in 1995 into five time intervals, and found that, despite the relatively small amplitude variations in the continuum flux, the Fe K$\\alpha$ line flux and profiles show dramatic variations: during the first time interval I-1, the iron line profile is broad and skewed, showing a strong peak at 6.4 keV, and a huge red wing extending to $\\sim$ 4.5 keV, similar to the time-averaged line profile; after several 10$^4$ s, the huge red wing becomes very weak (I-2); for I-3, except the strong peak at 6.4 keV, there is also a seeming peak at around 5.4 keV; the line profile of I-4 is similar to that of I-3, but much weaker; and finally, the line profile change back to single peak (I-5). The line equivalent widths of the five intervals are 375, 148, 329, 168 and 300 eV respectively, showing obvious variations. The rapid variabilities of Fe K line profiles discovered previously have suggested that the X-ray continuum flux is generated by magnetic flares above the accretion disk, and at any given time there are only a few flares moving around above the disk (Iwasawa et al. 1996, Fabian 1997, Wang et al. 1999b, Iwasawa et al. 1999, Wang et al. 2000). Flares with different location and different motion can give different fluorescence line profiles. Can the magnetic flare model also explain the observed rapid variabilities of Fe K$\\alpha$ line in NGC 4151? For I-1 and I-5, the X-ray continuum (in the flare model context) could be dominated by a single or several strong flares located very close to the center black hole, while the inner disk radius from the disk line fits (7.4 and 8.3 R$_g$) gives strong supports to this model. For I-2, the X-ray continuum seems to be dominated by flares far away from the center black hole (R$_i$ = 17 R$_g$). For I-3 and I-4, the X-ray continuum could be dominated by a single or several strong flares, which are $\\sim r_g$ above and coroating with the disk at around 14 R$_g$, with only the materials in the ring around 14 R$_g$ being illuminated. The large dramatic decreasing in the line EWs from I-3 to I-4 might be due to different degree of the anisotropy in the X-ray continuum, i.e., larger fraction of the X-ray continuum received by the disk in I-3 than in I-4. However, it requires some fine tune in order to keep the observed continuum more or less constant. Alternatively, Misra (2000) suggested that the changes in the disk structure may explain the observed line profile variations, but, a disk undergoing structure changes is also hard to keep the same radiative spectrum. Finally, the decreasing of the fluorescent yielding may be due to the increasing ionization in the disk skin. Under flare context, large change in the ionization structure of disk skin is possible even for quite moderate variations in the observed continuum flux if the flares have different heights If this is case, we would expect that the flare is lower in the I-4 than in I-3. This should also reflect in the line profile. However, the uncertainty in the disk parameters do not allow us to verify this. The observed properties strongly suggest that the rapid line variations may not be due the variations of X-ray continuum source. Instead, it may be due to changes in geometry or ionization level of the reflecting medium and at the same time, these changes should not affect the X-ray producing medium. If so, except for the variations of the X-ray continuum source, what can lead to the changes in geometry or ionization level of the reflecting medium? An absorption line at 5.8 keV is marginally detected in I-4, but not significant in other intervals. Note a variable absorption line at the same energy has been discovered by Nandra et al. (1999) in NGC 3516. Nandra et al. (1999) interpreted this line as the redshifted Fe K resonant absorption line. The large redshift is either caused by the infalling absorbing material or by the strong gravity in the vicinity of the black hole. Given the fact that the line is highly variable and the large redshift, the agreement of the redshifts in this two objects perhaps is not purely coincidence, but have some physical origin. If so, the gravitational redshift explanation would be favored." }, "0010/astro-ph0010530_arXiv.txt": { "abstract": "Relying on his photometry of galaxies, Holmberg (1958) found, more than 40 years ago, that the color indices of paired galaxies were closely correlated. Our deep broad-band BVRI CCD photometry of 45 (S+S) pairs from the Karachentsev (1972) catalogue (see Hern\\'andez Toledo \\& Puerari, this volume), and additional (B-V) color information from the literature (50 extra (S+S) pairs), help us to confirm the effect. This ``Holmberg Effect'' has long been remained unverified and not explained yet. ", "introduction": "Hints that galaxy interactions might play an important role in inducing multi-wavelength emission enhancement date back at least 40 years. One of the first quantitative indications came from Holmberg (1958), as a byproduct of his famous photometric survey of nearby galaxies. By comparing the photographic colors of paired galaxies, he found a significant correlation between the colors of pair components. This phenomenon has since been referred to as the ``Holmberg effect''. Although the physical explanation of the Holmberg effect is complex, it reflects in part a tendency for similar types of galaxies to form together (morphological concordance), a possible reflection of the role of local environment in determining galaxy morphology, but it can presumably also reflect mutually induced star formation (Kennicutt 1998). ", "conclusions": "" }, "0010/astro-ph0010195_arXiv.txt": { "abstract": "We test the methods developed in previous papers for inferring the intrinsic shapes of elliptical galaxies, using simulated objects from $N$-body experiments. The shapes of individual objects are correctly reproduced to within the statistical errors; close inspection of the results indicates a small systematic bias in the sense of underestimating both the triaxiality and short-to-long axis ratio. We also test the estimation of parent shape distributions, using samples of independent, randomly oriented objects. The best results are statistically accurate, but on average the parent distributions are again slightly biased in the same sense. Using the posterior probability densities for the individual objects, we estimate the magnitude of the bias to be $\\lesssim 0.1$ in both shape parameters. The results support the continued use of these methods on real systems. ", "introduction": "} Observational estimates of the true three-dimensional shapes of elliptical galaxies serve as diagnostics of the physics of galaxy formation and evolution. Numerical simulations of protogalactic collapse (Dubinski \\& Carlberg \\markcite{DC91}1991), mergers (Weil \\& Hernquist \\markcite{WH96}1996, Barnes \\& Hernquist \\markcite{BH98}1998), and black-hole growth (Merritt \\& Fridman \\markcite{MF98}1996, Merritt \\& Quinlan \\markcite{MQ98}1998) all suggest that dark halos {\\it should\\/} tend toward axisymmetry with increasing dissipation or central mass concentrations. But whether this has, in fact, happened is difficult to establish. The effort to uncover the intrinsic shape distribution has a long history (reviewed by Statler \\markcite{Sta95}1995, updated by Bak \\& Statler \\markcite{BS00}2000, hereafter BS). Photometric methods, introduced by Hubble \\markcite{Hub26}(1926) and revisited by many authors (most recently Ryden \\markcite{Ryd92}1992, \\markcite{Ryd96}1996; Fasano \\markcite{Fas95}1995), have been effective in constraining the distribution of flattenings, but not in establishing the relative numbers of triaxial and axisymmetric systems. For constraining triaxiality, dynamical information is necessary, with both gas (Bertola et al. \\markcite{Ber91}1991) and stellar (Binney \\markcite{Bin85}1985, Franx et al. \\markcite{FIZ91}1991, Tenjes et al. \\markcite{Ten93}1993) kinematic data having been used for this purpose. Earlier papers in this series (Statler \\markcite{S94a}1994a, Statler \\& Fry \\markcite{SF94}1994, Statler \\markcite{S94b}1994b, hereafter Papers I, II, and III respectively) presented a method for statistically estimating the intrinsic shapes of elliptical galaxies from observations of their radial velocity fields and surface brightness distributions. Paper III included a test of the method on the end product of an $N$-body collapse simulation (Dubinski \\markcite{D92}1992). Using simulated observations of the object along a single line of sight, the method could recover the true shape only to within a $2\\sigma$ error region; this result was attributed to an unfortunate line of sight. Since then, the method has profitably been applied to real systems, to derive both the shapes of individual galaxies (NGC 3379, Statler \\markcite{S94c}1994c, Statler \\markcite{S00}2001, NGC 1700, Statler et al.\\ \\markcite{SDS99}1999, hereafter SDS) and the parent shape distribution of a sample (Davies \\& Birkinshaw \\markcite{DB88}1998) of radio ellipticals (\\markcite{BS00}BS). Nonetheless, there remains a lingering impression that the performance on realistic simulated systems is ``disappointing'' (Binney \\& Merrifield \\markcite{BM98}1998). In this paper we present more extensive tests, using a homogeneous set of simulated elliptical galaxies from group-merger experiments (Weil \\& Hernquist \\markcite{WH94}1994, \\markcite{WH96}1996, hereafter WH94 and WH96, collectively WH). We demonstrate that the method performs well, both in estimating the shapes of individual objects and in deriving the shape distribution of small samples. The presentation is ordered as follows: Section 2 briefly describes the test objects and the methods employed, with references to earlier papers where complete expositions can be found. Section 3 describes the tests performed and their results. Section 4 deals with residual systematic bias and its possible origin, and Section 5 sums up. ", "conclusions": "} We would like to understand the origin of the small systematic bias in our results. One possibility is that the small samples used in \\S\\ \\ref{s.samples} could be biased in orientation, but this is not the case. Figure \\ref{f.orientations} shows the distribution of the lines of sight, relative to the principal axes of the observed objects, for each sample. There is no obvious clustering of views. The isotropy of the distribution can be quantified using $|Q_1|$, the largest (in absolute magnitude) eigenvalue of the quadrupole moment tensor, normalized to the number of objects. A Monte Carlo simulation of $10^5$ six-object samples gives a skewed distribution with a mean of $0.44$, a mode of $0.36$, and a standard deviation of $0.22$. The $|Q_1|$ values for Samples 1, 2, and 3 are $0.35$, $0.42$, and $0.24$ respectively, so the samples are neither abnormally anisotropic nor abnormally isotropic. \\begin{figure}[t]{\\epsfxsize=4.5in\\hfil\\epsfbox{f4.eps}} \\caption{\\footnotesize Distribution of lines of sight for the observed objects in ($a$) Sample 1, ($b$) Sample 2, ($c$) Sample 3. In each panel the origin corresponds to views down the short axis, the top edge to views down the long axis, and the right edge to views down the middle axis. For these models views are uniquely defined over only 1 octant, which is shown in equal-area projection with dotted lines at $30\\arcdeg$ intervals. \\label{f.orientations}} \\end{figure} At any rate, bias in the 6-object samples, even if present, would not account for the similar systematic error seen in the single-object tests of \\S\\ \\ref{s.single}. Unfortunately, we see no obvious reason why the method should tend to underestimate the triaxiality and overestimate the flattening. But we can speculate on three possibilities: \\begin{enumerate} \\item {\\em Overly cautious use of the kinematic data.\\/} The observed rotation curves are averaged over a range of radii where they are well-behaved. By indulging a natural tendency to exclude radii where the data become ``weird,'' we may have reduced the signature of the very kinematic asymmetries that are the hallmarks of triaxiality. \\item {\\em Coarse dynamical grid.\\/} \\markcite{SDS99}SDS and \\markcite{BS00}BS emphasized the importance of the ``disklike'' or ``spheroid-like'' character of the internal velocity field to the inferred shape. The model grid used here includes only the extreme cases on the assumption that they should bracket the correct result. However, we occasionally encounter situations where shapes inferred from intermediate models are more triaxial than those from either extreme (Statler \\markcite{S00}2001). \\item {\\em Figure rotation.\\/} Pattern speeds for the merger remnants have not been measured. An addition of a solid-body component to the streaming velocity field of a triaxial object can, in principle, make it appear more axisymmetric than it actually is. This is a long-standing unresolved issue, and there are still no dynamical models for ellipsoidal systems that include figure rotation as an adjustable parameter. \\end{enumerate} } We have tested the methods developed in previous papers for inferring the intrinsic shapes of elliptical galaxies, using a set of six simulated objects formed by group mergers of disk/halo systems. We have modeled two of the objects over a uniform grid of orientations, and find that their true shapes are recovered to within the statistical errors. A more stringent test, using the prior knowledge that the same object is being observed repeatedly, indicates a small systematic bias in the results, in that both triaxiality and short-to-long axis ratio are underestimated by roughly $0.1$. We have also constructed three simulated samples of 6 objects, and estimate the parent shape distribution in each case. The result for one of the samples is statistically unimpeachable, with the correct fractions of the sample falling within the 68\\% and 95\\% HPD regions. The other two samples give results that are worse by an amount consistent with the above bias. Including prior knowledge of the internal dynamics of the objects improves the results for all the samples. Examination of the posterior shape estimates for the individual objects in the samples again suggests a systematic bias $\\lesssim 0.1$ in triaxiality and flattening, in the same sense found in the earlier test. The source of this bias is not understood, but may be related to the handling of the observational data, the coarse grid of dynamical parameters in the models, or figure rotation of the test objects. As a whole, these results support the use of our methods to understand the nature of real elliptical galaxies." }, "0010/astro-ph0010476_arXiv.txt": { "abstract": "The BeppoSAX observations of the bright Narrow--Line Seyfert 1 galaxy Ark~564 are presented along with a high quality optical spectrum taken at the 1.5m telescope at La Silla. The 0.1--10 keV X--ray spectrum is characterized by a strong soft component which is best described by blackbody--like emission with a temperature of $\\sim$ 160 eV. At higher energies a steep ($\\Gamma \\simeq$ 2.4) power--law tail is present. There is evidence of an ionized reflector in the form of an iron line and edge. We do not find significant evidence of soft X--ray features if the spectrum is modelled with a two component continuum. The optical and X--ray spectral properties support the hypothesis of a high accretion rate onto a low mass black hole. ", "introduction": "In recent years sensitive ROSAT and ASCA observations have revealed that a class of objects classified on the basis of their optical spectra as Narrow--Line Seyfert 1 galaxies (hereinafter NLS1) exhibits peculiar and rather extreme properties in the X--ray band. Large amplitude X-ray variability with timescales as fast as a few minutes (Boller et al. 1997) and extremely soft X--ray spectra in the ROSAT band are common among these objects (Boller et al. 1996). ASCA observations have extended the ROSAT results to higher energies. On average, NLS1 galaxies have steeper X--ray spectra (Brandt et al. 1997) and higher variability amplitude variance (Fiore et al. 1998; Leighly 1999a; Turner et al. 1999a) compared to Seyfert 1s with broad optical lines. The anti--correlation between the X--ray spectral slope and the H$\\beta$ FWHM is now a well--established observational property tested across a wide range of X--ray luminosities (Laor et al. 1997), suggesting that NLS1 represent an extreme of Seyfert activity, possibly linked to an extreme value of a fundamental physical parameter. The ASCA observation of REJ 1034+393 can be considered a breakthrough in the understanding of NLS1, because it was the first object of this class observed above 2 keV. The 0.5--10 keV spectrum is dominated by a strong soft component below $\\sim$ 2 keV; a steep ($\\Gamma \\simeq$ 2.6) power law tail at higher energies and an ionized iron K$\\alpha$ line at 6.7 keV (although detected only at the 2$\\sigma$ level) are also present (Pounds et al. 1995). Ionized iron lines have been discovered in a few other objects observed by BeppoSAX (Comastri et al. 1998, paper I) and ASCA (Turner et al. 1998; Leighly 1999b), suggesting that the fundamental parameter could be a particularly high accretion rate (Pounds et al. 1995; Laor et al. 1997). We have started an observational programme to study a sizeable sample of NLS1 with BeppoSAX with the aim of investigating the broad band X--ray spectral and variability properties of these objects. We are taking advantage of the increased effective area at high energies, especially above 5 keV, where the MECS detectors provide the best opportunity to investigate the iron line properties. Arakelian 564 ($z$ = 0.0247, $V$ = 14.16) is an ideal candidate for such an investigation being the brightest NLS1 in the 2--10 keV band. Previous ROSAT and ASCA observations revealed a complex X--ray spectrum with a few absorption/emission features around 1 keV (Brandt et al. 1994; Leighly 1999b; Vaughan et al. 1999a; Turner et al. 1999b, hereinafter TGN99). More recently, an RXTE/PCA observation (Vaughan et al. 1999b, hereinafter V99) provided the first spectrum of this object up to 20 keV. The most remarkable result is the presence of a deep ionized edge at $\\sim$ 8.5 keV interpreted as reflection from an ionized disc. In the present paper we report the results of 3 BeppoSAX observations of Ark 564 together with a high--quality optical spectrum obtained at La Silla one year before the first X--ray observation. ", "conclusions": "We have analysed BeppoSAX data and a high-quality optical spectrum of the bright Narrow-Line Seyfert~1 galaxy Ark~564. Our main results are the following: $\\bullet$ The optical emission--line intensities and widths clearly confirm the NLS1 nature of Ark~564. The Fe~{\\sc ii} multiplets are not particularly strong and are similar to those of normal Seyfert 1 galaxies. $\\bullet$ The evidence for soft X--ray spectral features is marginal if the strong and possibly complex soft X--ray excess is modelled with a blackbody or a multicolour accretion disc spectrum. $\\bullet$ The spectrum hardens at high energies though the power law photon index is rather steep and significantly steeper than the average value measured in broad line Seyfert 1 galaxies. $\\bullet$ The high--energy iron K$\\alpha$ spectral features (the $\\sim$ 6.8 keV line and the $\\sim$ 9.5 keV edge) strongly suggest reflection from a highly ionized, optically thick accretion disc. $\\bullet$ The optical to X--ray spectral energy distribution peaks in the soft X--ray band with a behaviour similar to that observed in RE J1034+393. A rapidly accreting, low mass black hole provides the most likely interpretation of the broad band observations." }, "0010/astro-ph0010126_arXiv.txt": { "abstract": "We compute precise predictions for the two-point correlation function of local maxima (or minima) in the temperature of the microwave background, under the assumption that it is a random gaussian field. For a given power spectrum and peak threshold there are no adjustable parameters, and since this analysis does not make the small-angle approximation of \\scite{HS99}, it is essentially complete. We find oscillatory features which are absent in the temperature autocorrelation function, and we also find that the small-angle approximation to the peak-peak correlation function is accurate to better than 0.01 on all scales. These high-precision predictions can form the basis of a sensitive test of the gaussian hypothesis with upcoming all-sky microwave background experiments MAP and Planck, affording a thorough test of the inflationary theory of the early Universe. To illustrate the effectiveness of the technique, we apply it to simulated maps of the microwave sky arising from the cosmic string model of structure formation, and compare with the bispectrum as a non-gaussian discriminant. We also show how peak statistics can be a valuable tool in assessing and statistically removing contamination of the map by foreground point sources. ", "introduction": "The cosmic microwave background radiation (CMB) presents an ideal opportunity to test theories of the early Universe. At the time of last scattering, the Universe is a relatively straightforward, almost uniform, mixture of photons, baryons, electrons and dark matter. The physics is well-understood, and free from the very complicated effects which make interpretation of the present-day matter distribution more complicated. The microwave background thus offers the possibility of accurately testing models of structure formation. A generic test can readily be made between two classes of structure-formation models, based on inflation and cosmic defects respectively. There are several ways to do this; the power spectrum itself is a useful discriminant of specific models. We concentrate here on a generic test: most inflationary models predict that the microwave background temperature map will be very close to a random gaussian field, whereas generically defect models predict a non-gaussian temperature map. It turns out that testing the gaussian nature of the initial fluctuations is easier through analysis of CMB fluctuations than large-scale structure \\cite{VWHK00}, although tests based on number densities of high-redshift objects may also be useful (\\pcite{Robinson00},\\pcite{MVJ00}). Current evidence from Boomerang \\cite{Boomerang} and MAXIMA (\\pcite{MAXIMA1},\\pcite{MAXIMA2}) favours inflation models, since the power spectrum is acceptable for certain combinations of cosmological parameters. Indeed, the major scientific goal of these and future experiments such as the {\\em Microwave Anisotropy Probe} (MAP) and {\\em Planck Surveyor} \\cite{PhaseA}, is to derive cosmological parameters from the power spectrum. To make this interpretation requires that the temperature map is created by inflation or some similar process, not by defects, and that the map is not seriously contaminated by foregrounds. In both of these areas, the statistics of peaks can be a valuable tool. The process is quite straightforward: given a power spectrum, the statistical properties of peaks of a gaussian field are fully determined - there are no free parameters. If the peaks are not consistent with the predictions, then either the CMB temperature map is not gaussian, or it is significantly contaminated by foregrounds, or both. In either of these cases, the derived cosmological parameters from the power spectrum will be suspect. In this paper, we compute the predictions for the correlation function of local maxima (and minima) for a gaussian field. The paper generalises the work of \\scite{HS99} in dropping the small-angle approximation: the results of this paper can be used for all valid separations on the sky. There are several ways to test the gaussian hypothesis, such as the three-point function (e.g. \\pcite{Hinshaw94}, \\pcite{FRS93}, \\pcite{LS93}, \\pcite{GLMM94}), the genus and Euler-Poincar\\'e statistic (\\pcite{Coles88}, \\pcite{Gott90}, \\pcite{Luo94b}, \\pcite{Smoot94}), the bispectrum (\\pcite{Luo94}, \\pcite{Hea98}, \\pcite{Ferreira98}), studies of tensor modes in the CMB \\cite{CCT94}, excursion set properties (\\pcite{Barreiro98}, \\pcite{Barreiro2000}), peak statistics (\\pcite{BE87}, \\pcite{Kogut95}, \\pcite{Kogut96}, \\pcite{Barreiro97}) and wavelet analyses (e.g. \\pcite{MHL00}, \\pcite{AF99}, \\pcite{FA99}). One advantage which the method presented here has is the possibility of assessing and removing contamination by foreground point sources. We return to this in the discussion. Non-gaussian signals have been reported for the COBE map by \\scite{Ferreira98} (see also \\pcite{Pando98}, \\pcite{KJ98}, \\pcite{BT99}, \\pcite{MHL00}, \\pcite{Mag2000}) . If this nongaussian signal is really present in the microwave background map, and not the result of some artefact \\cite{Banday2000}, then it would be a severe challenge to inflation models, as it is many orders of magnitude larger than expected (e.g. \\pcite{VWHK00} and references therein). ", "conclusions": "We have presented calculations of the exact correlation function of peaks in a random gaussian field defined on the surface of a sphere. No small-angle approximation is made, so the method is an advance on the flat-sky computations of \\scite{HS99} and now effectively complete. The formalism allows very accurate theoretical predictions of the peak-peak correlation function for temperature fluctuations in the microwave background, which is the application considered here. We envisage the main use of this method being as a sensitive test of the gaussian hypothesis. Since inflationary models generically predict a temperature field which is very close to gaussian, this is a consistency test for inflation. Other structure formation models, based for example on strings, predict non-gaussian temperature maps. Although the visual appearance of string maps is evidently non-gaussian, it is not necessarily easy to find statistics which will unambiguously distinguish them from gaussian fields. To illustrate this point, we have analysed 12.5-degree square simulated maps of string models, using the bispectrum and the peak-peak correlation function as distinguishing statistics. We find that, while cosmic variance in the bispectrum makes it difficult to use on a small patch of sky, the peak-peak correlation function clearly rules out a gaussian map. In practice, maps of the microwave background will be contaminated at some level by point sources, amongst other things. Peak statistics may be useful in assessing this contribution. The most straightforward example is that an uncontaminated map has the same average number density of maxima and minima; a significant excess of maxima would be indicative of contamination. Unfortunately the theory of peaks is not able to tell us the {\\em distribution} of the number of maxima or minima within a finite sky (only its mean), but it is a straightforward matter to determine the distribution by monte carlo realisations. One can attempt to go further than this, by removing statistically the contribution from the point sources, provided one knows from other observations what their correlation function is. Assuming the point sources are uncorrelated with the microwave background peaks, the correlation function of the combined map is simply a weighted mean of the two. The point sources will contribute to the power spectrum; one can vary the assumed contribution from point sources and modify the power spectrum and the derived microwave background peak correlation function accordingly. If consistency can be achieved, one will be confident both of the gaussian nature of the microwave background, and the level of point source contamination. {\\bf Acknowledgments} We are grateful to Francois Bouchet for providing the simulated catalogues, and to Ravi Sheth for useful discussions. Computations were made partly using Starlink facilities." }, "0010/astro-ph0010310_arXiv.txt": { "abstract": "We present a first emission line analysis of a high resolution X-ray spectrum of the stellar wind of $\\theta^1$ Ori C obtained with the High Energy Transmission grating Spectrometer onboard the Chandra X-ray Observatory. The spectra are resolved into a large number of emission lines from H- and He-like O, Ne, Mg, Si, S, Ar and Fe ions. The He-like Fe XXV and Li-like Fe XXIV appear quite strong indicating very hot emitting regions. From H/He flux ratios, as well as from Fe He/Li emission measure ratios we deduce temperatures ranging from 0.5 to 6.1$\\times 10^7$ K. The He-triplets are very sensitive to density as well. At these temperatures the relative strengths of the intercombination and forbidden lines indicate electron densities well above 10$^{12}$ cm$^{-3}$. The lines appear significantly broadened from which we deduce a mean velocity of 770 km s$^{-1}$ with a spread between 400 and 2000 km s$^{-1}$. Along with results of the deduced emission measure we conclude that the X-ray emission could originate in dense and hot regions with a characteristic size of less then 4$\\times 10^{10}$ cm. ", "introduction": "X-ray spectra from stellar winds have been quite difficult to interpret, since the first observations with \\ein (Seward et al. 1979). The spectral resolution from previous X-ray missions was insufficient for resolving any emission and absorption lines. Models that describe the mass loss from a hot luminous star successfully in the UV (Pauldrach et al. 1994, Lamers et al. 1999) could never correctly predict observed X-ray fluxes. A model involving a possible hot corona seems unlikely due to the lack evidence for absorption at the 20.6 \\AA~ K-shell ionization edge from oxygen (Cassinelli $\\&$ Swank 1983); additionally no coronal line emission (i.e. Fe XIV at 530 nm) was found in optical spectra of $\\eta$ Ori and $\\kappa$ Ori (Nordsieck et al. 1981) nor in the protoptype star for stellar winds $\\zeta$ Pup (Baade $\\&$ Lucy (1987). X-ray emission from shocks emerging in instabilities within a radiatively driven wind forming a forward shock was proposed by Lucy $\\&$ White (1980) and Lucy (1982). However, reverse shocks that decelerate material as it rams into dense shells ahead seem more likely (Owocki et al. 1988). These models predict X-rays up to a temperature of about kT$\\sim$0.5 keV. Feldmeier et al. (1997) introduced turbulent perturbations into their calculations which ultimatively allowed for higher temperatures. Several recent observations with the 0.1 to 10.0 keV \\asca bandpass, i.e. of $\\tau$ Sco (Cohen et al. 1997) and $\\eta$ Car (Tsuboi et al. 1997), indicated that there is a hard component to these spectra at $>$ 2 keV. Although these two examples tend to resemble quite extreme cases of massive stars, \\asca observations of other, less massive O-stars (Corcoran et al. 1994), and the Orion Trapezium (Yamauchi et al. 1996) indicated the existence of a similar hard component in their spectra as well. In the most recent analysis of \\chandra CCD spectra of the Orion Trapezium, Schulz et al. (2000) were able to resolve the entire Trapezium into individual sources. Identified O- and B-stars showed a soft component at $\\sim$ 0.8 keV as observed with \\ros; the very early spectral types required an additional hard component of temperatures above 2 keV as well. In this letter we present first results from an observation of $\\theta^1$ Ori C with the High Energy Transmission Grating Spectrometer (HETGS, Canizares et al. 2000, in preparation) onboard \\chandra (Weisskopf et al. 1996). We present an X-ray line list of the brightest emission lines detected and investigate emission line properties to determine the range of temperatures and densities of the line emitting regions. Some of the results will have immediate impact constraining models of stellar winds. \\centerline{\\epsfxsize=7.5cm\\epsfbox{f1.ps}} \\figcaption{HETGS focal plane image of the Orion Trapezium Cluster \\label{focal}} ", "conclusions": "A large number of emission lines were detected with a signal to noise ratio of greater than 5$\\sigma$. These lines are from H- and He-like ion species of Fe, S, Si, Mg, Ne, and O; Ca and Ar are not detected. We do not observe any residual line shifts we could attribute to the source. The X-ray lines appear significantly broadened implying a mean shock velocity of 770 km s$^{-1}$. A similar shock velocity of 500 km s$^{-1}$ has been determined from \\ros fits of $\\zeta$ Pup spectra (Hillier et al. 1993). Within the statistical uncertainities and the current state of the analysis, the line profiles appear symmetric showing blue and red-shifts of equal proportions. We therefore conclude that the wind emission is quite isotropic. Our estimates from the VEM indicates that the emission volume is equivalent to a volume of well within one stellar radius. However, Simulations by Feldmeier et al. (1997) indicated that the X-ray emission must extend far out into the wind. Estimating the distance of the X-ray producing shell collision we use the $\\beta$ velocity law (with $\\beta$= 0.88) and a terminal velocity of $\\sim$ 1000 km s$^{-1}$ (Prinja et al. 1990) to determine the distance where the wind reaches 770 km s$^{-1}$, which gives about 4 stellar radii. In this respect it is plausible that the X-ray emitting regions appear in very thin dense shells as described by Feldmeier et al. (1997). Their model, however, predicts strong variability of the X-ray emission, which, at least in terms of total X-ray flux, is not observed during our observations. Whatever the actual emission geometry turns out to be, the emitting regions of $\\theta^1$ Ori C appear well confined, hot, and dense with a temperatures ranging between 0.05 and 6 $\\times 10^7$ K (0.06 to 4.5 keV) and densities above $10^{12}$ cm$^{-3}$. CCD spectra of $\\theta^1$ Ori C already indicated that the X-ray emission must have a hot component with a temperature of kT$\\sim$3 keV (Yamauchi et al. 1997, Schulz et al. 2000) or equivalent to $\\sim$ 4$\\times 10^7$ K. A broadband study ranging from UV to \\asca spectra of the B0 V star $\\tau$ Sco Cohen (1996) also suggested that wide range of temperatures up to $10^7$ K is necessary to describe X-ray and UV emission. It is quite hard to explain such high temperature with the line driven shock instability models (Lucy $\\&$ White 1980, Lucy 1982), which predict temperatures more on the order of 5$\\times 10^6$ K. The spectrum in Figure 2 also shows that the hard part of the spectrum cannot be explained by inverse Compton scattering of UV photons (Chen $\\&$ White 1991), since it shows emission lines with similar characteristics than the low energy lines. The approach by Feldmeier et al. (1997) seems to provide promising ingredients to describe the co-existence of low and high temperatures at relatively high densities. A most recent analysis by Kahn et al. (2000) of XMM RGS spectra of $\\zeta$ Pup also resulted in low f/i line ratios and thus high density values, but it was suggested that a high UV radiation field could destroy the He-like forbidden line leading to an overestimation of the density. It still remains to be explained whether the high temperatures in $\\theta^1$ Ori C and other hot candidates are the exception or the rule. Clearly, we need more high resolution X-ray observations of massive O-Stars in order classify the X-rays in terms of their line emission properties. These properties include temperature ranges, densities, emission volumes and ionization balance." }, "0010/astro-ph0010583_arXiv.txt": { "abstract": "A Virtual Observatory (VO) will enable transparent and efficient access, search, retrieval, and visualization of data across multiple data repositories, which are generally heterogeneous and distributed. Aspects of data mining that apply to a variety of science user scenarios with a VO are reviewed. ", "introduction": "What is data mining and why is applicable to scientific research? Data mining is defined as {\\it{an information extraction activity whose goal is to discover hidden facts contained in databases}}. Data mining has taken the business community by storm and there is consequently now a vast array of resources and research techniques available for exploitation by the scientific communities. It is useful therefore to examine a categorization of data mining thrusts and their sub-components, since these are likewise applicable to the scientific exploration of large astronomical databases. Data mining is used to find patterns and relationships in data by using sophisticated techniques to build models -- abstract representations of reality. A good model is a useful guide to understanding that reality and to making decisions. There are two main types of data mining models: {\\it{descriptive}} and {\\it{predictive}}. {\\it{Descriptive}} models describe patterns in data and are generally used to create meaningful subgroups or clusters. {\\it{Predictive}} models are used to forecast explicit values, based upon patterns determined from known results. There is another differentiation of data mining into two categories that we find particularly appropriate to knowledge discovery in large astronomical databases: {\\it{event-based mining}} and {\\it{relationship-based mining}}. At the risk of trivializing some fairly sophisticated techniques, we classify event-based mining scenarios into four orthogonal categories: \\begin{itemize} \\item Known events / known algorithms -- use existing physical models ({\\it{descriptive models}}) to locate known phenomena of interest either spatially or temporally within a large database. \\item Known events / unknown algorithms -- use pattern recognition and clustering properties of data to discover new observational (in our case, astrophysical) relationships among known phenomena. \\item Unknown events / known algorithms -- use expected physical relationships ({\\it{predictive models}}) among observational parameters of astrophysical phenomena to predict the presence of previously unseen events within a large complex database. \\item Unknown events / unknown algorithms -- use thresholds to identify transient or otherwise unique (``one-of-a-kind'') events and therefore to discover new phenomena. \\end{itemize} \\noindent For relationship-based mining, we identify three classes of association-driven scenarios that would find application in astronomical research: \\begin{itemize} \\item Spatial associations -- identify events (astronomical objects) at the same location in the sky. \\item Temporal associations -- identify events occurring during the same or related periods of time. \\item Coincidence associations -- use clustering techniques to identify events that are co-located within a multi-dimensional parameter space. \\end{itemize} From this discussion, we thus derive a reduced set of science requirements for data mining that correspond to the following exploratory approaches to mining large databases\\thinspace : {\\it{Object Cross-Identification}}, {\\it{Object Cross-Correlation}}, {\\it{Nearest-Neighbor Identification}}, and {\\it{Systematic Data Exploration}}. (a)~``Object cross-identification'' refers to the classical problem of connecting the source list in one catalog (or observation database) to the source list in another, in order to derive new astrophysical understanding of the cross-identified objects (e.g., gamma-ray burst counterparts). (b)~``Object cross-correlation'' refers to the application of ``what if'' scenarios to the full suite of parameters in a database (e.g., identify distant galaxies as $U$-band dropouts in a color-color scatter plot from the HDF survey). (c)``Nearest-neighbor identification'' refers to the general application of clustering algorithms in multi-dimensional parameter space (e.g., finding the closest known population of young stars -- in the TW Hydrae association -- through their similar kinematics, X-ray emission, $H\\alpha$, and Li abundance). (d)~``Systematic data exploration'' refers to the application of the broad range of event-based and relationship-based queries to a database in the hope of making a serendipitous discovery of new objects or a new class of objects (e.g., finding new types of variable stars, such as ``bumpers'', in the MACHO database). ", "conclusions": "" }, "0010/astro-ph0010060_arXiv.txt": { "abstract": "Results are presented from a long-term astrometry program on PSR B0919+06 using the NRAO Very Long Baseline Array. With ten observations (seven epochs) between 1994--2000, we measure a proper motion $\\mu_{\\alpha} = 18.35 \\pm 0.06$ mas yr$^{-1}$, $\\mu_{\\delta} = 86.56 \\pm 0.12$ mas yr$^{-1}$, and a parallax $\\pi = 0.83 \\pm 0.13$ mas (68\\% confidence intervals). This yields a pulsar distance of $1.21 \\pm 0.19$ kpc, making PSR B0919+06 the farthest pulsar for which a trigonometric parallax has been obtained, and the implied pulsar transverse speed is $505 \\pm 80$ \\kms. Combining the distance estimate with interstellar scintillation data spanning 20 years, we infer the existence of a patchy or clumpy scattering screen along the line of sight in addition to the distributed electron density predicted by models for the Galaxy, and constrain the location of this scattering region to within $\\sim 250$ parsecs of the Sun. Comparison with the lines of sight towards other pulsars in the same quadrant of the Galaxy permits refinement of our knowledge of the local interstellar matter in this direction. \\\\ ", "introduction": "Since the discovery of pulsars, it has been recognized that the dispersion and scattering of pulsar signals provide unique information about the intervening medium. The distance to most pulsars is estimated using the observed dispersion measure (DM) and a model for the Galactic electron density distribution \\citep[e.g.][hereafter TC93]{TC93}. Where available, a model-independent distance from annual trigonometric parallax provides crucial calibration information for this model, as well as allowing the absolute luminosity of the pulsar to be derived. Pulsar astrometry also provides the observational evidence required to investigate several other questions. Proper motion measurements (especially in conjunction with reliable distance estimates) allow verification of pulsar--supernova remnant associations \\citep[e.g.][]{K98}. The inferred speeds of pulsars constrain the minimum asymmetry in supernova core-collapse processes, or other sources of kick velocities. Pulsar population statistics, selection-effect biases, and planetary--extragalactic reference frame ties have also been addressed by astrometry \\citep[e.g.,][]{LL94,CC98,BCR+96}. At present, there are only a handful of model-independent distances to pulsars. In a recent summary, \\citet{TBM+99} list nine such objects. In \\S\\ref{Sec:VLBI} of this paper, we present results from a long-term astrometry program on PSR B0919+06 using the NRAO Very Long Baseline Array (VLBA). We measure the proper motion and trigonometric parallax for this pulsar, and infer a distance of $\\sim 1.2$ kpc and a transverse speed of $\\sim 500$ \\kms, comparable to the mean population speed. PSR B0919+06 has a DM of 27.31 pc cm$^{-3}$ \\citep{PW92}, yielding a mean electron density of 0.023 cm$^{-3}$ for the interstellar medium (ISM) along this line of sight. However, interstellar scintillation (ISS) and scattering of the pulsar signals can be used to extract much more information about the line of sight than simply this mean electron density. In \\S\\ref{Sec:ISS}, we review the necessary formalism, and analyze the published data spanning $\\sim 20$ years for this pulsar. Our analysis combines data from very long baseline interferometry (VLBI) and ISS, in order to infer the distribution of scattering material along the line of sight. We find the need for extra scattering material in addition to the TC93 distribution; with the assumption that this material is present in the form of a screen or clump with variable scattering strength, we constrain the scren location to the local interstellar medium, within $\\sim 250$ parsecs of the Sun. This analysis, along with DM and distance measurements for other pulsars in the third quadrant of our Galaxy, allows refinement of our knowledge of the local electron density distribution. Besides being a situation of interest in its own right, this analysis also serves as a model for the hybrid VLBI--ISS technique applied to the local ISM. ", "conclusions": "Parallax distances to pulsars have the prospect of resolving several outstanding questions, both about the pulsar population and about the intervening ISM. In this work, we have presented VLBA astrometry on PSR B0919+06, using phase-referenced observations with an in-beam calibrator only 12\\arcmin\\ from the pulsar. The use of the in-beam source reduces ionospheric effects and other astrometric errors to the 0.1 mas level and allows sub-\\mas\\ astrometry. With ten observations (seven distinct epochs) distributed over seven years, we derive a proper motion $\\mu_{\\alpha} = 18.35 \\pm 0.06$ mas yr$^{-1}$, $\\mu_{\\delta} = 86.56 \\pm 0.12$ mas yr$^{-1}$, and measure a parallax $\\pi = 0.83 \\pm 0.13$ mas (68\\% confidence intervals). This result implies a transverse speed of $505 \\pm 80$ \\kms\\ at a distance $D = 1.21 \\pm 0.19$ kpc, making PSR B0919+06 one of the farthest objects for which a trigonometric parallax has been obtained. The use of an in-beam calibrator with the VLBA at 1.5~GHz can provide trigonometric parallaxes to twice this distance, or even further (to 10~kpc) with the use of more sensitive antennas and higher frequencies. For the measured distance of 1.2~kpc and a DM of 27.31 pc~cm$^{-3}$ \\citep{PW92}, the derived mean electron density towards B0919+06 is 0.023 cm$^{-3}$. This line of sight is along the same Galactic longitude as the superbubble proposed by \\citet{H98}, as well as clouds of ionized gas in the local neighbourhood \\citep{DG98}, though at a higher Galactic latitude. We use scintillation data spanning 20 years in a hybrid analysis, equating the derived ISS and interferometric speeds of the pulsar in order to constrain the distribution of scattering material along the LOS. We find that excess scattering material is required compared to the standard TC93 model, and assuming that this excess material is confined to a thin screen or clump, we constrain its location to within $\\sim 250$~pc of the Sun. Comparison with the neighboring lines of sight to pulsars B0950+08, J1024$-$0719, and B0823+06 indicates that this result is consistent with a turbulent interface region between the Local Bubble and the GSH~238+00+09 superbubble, which possibly contains clumps of scale size $< 30$~pc. We note that using an updated model of the Galactic electron density distribution (Cordes {\\&} Lazio, in preparation) will not materially alter these conclusions." }, "0010/astro-ph0010256_arXiv.txt": { "abstract": "We propose and implement a fast, universally applicable method for extracting the angular power spectrum ${\\cal C}_{\\ell}$ from CMB temperature maps by first estimating the correlation function $ \\xi(\\theta)$. Our procedure recovers the ${\\cal C}_{\\ell}$'s using $N^2$ (but potentially $N \\log N$), operations, where $N$ is the number of pixels. This is in contrast with standard maximum likelihood techniques which require $N^3$ operations. Our method makes no special assumptions about the map, unlike present fast techniques which rely on symmetries of the underlying noise matrix, sky coverage, scanning strategy, and geometry. This enables for the first time the analysis of megapixel maps without symmetries. The key element of our technique is the accurate multipole decomposition of $\\xi (\\theta)$. The ${\\cal C}_{\\ell}$ error bars and cross-correlations are found by a Monte-Carlo approach. We applied our technique to a large number of simulated maps with Boomerang sky coverage in $81000$ pixels. We used a diagonal noise matrix, with approximately the same amplitude as Boomerang. These studies demonstrate that our technique provides an unbiased estimator of the ${\\cal C}_{\\ell}$'s. Even though our method is approximate, the error bars obtained are nearly optimal, and converged only after few tens of Monte-Carlo realizations. Our method is directly applicable for the non-diagonal noise matrix. This, and other generalizations, such as minimum variance weighting schemes, polarization, and higher order statistics are also discussed. ", "introduction": "Future missions of measuring the Cosmic Microwave Background (CMB) fluctuations will revolutionize our knowledge of cosmology. They will either confirm or refute the basic Big Bang paradigm, will reveal the values of most cosmological parameters within a few percent (e.g., Spergel 1994, Knox 1995, Hinshaw, Bennett, \\& Kogut 1995, Jungman \\etal 1996, Zaldarriaga, Spergel, \\& Seljak 1997, Bond, Efstathiou, \\& Tegmark 1997). The large number of pixels contained in current and future experiments enable these exciting developments, but at the same time, present unprecedented challenges the data analysis. The mainstream maximum likelihood techniques are already pushed to their limits by the largest existing CMB maps, but they are clearly inadequate for future megapixel surveys. Therefore the most important near term task for CMB research is to find techniques which could perform the required analyses with realistic resources, thus fulfill the promise of the high precision experiments. This Letter proposes a fast, universally applicable method for the estimation of $C_l$'s from any CMB pixel map, based on estimating correlation functions. The recovered errorbars are at most $10\\%$ larger than the theoretically smallest possible ones imposed by cosmic variance and noise. Optimal methods could further decrease these errorbars only slightly, and at an unrealistic cost; thus they represent a case of diminishing returns. Our procedure scales as $N^2$, and more sophisticated algorithms will improve this to $N \\log N$. This essentially solves the problem for two major upcoming space missions, MAP (Microwave Anisotropy Probe) and the Planck, and at the same time allows more detailed analyses of current and future balloon-borne and ground based experiments. The short analysis turn around time enables Monte-Carlo (MC) studies of systematics, foregrounds, errors, underlying models etc. These are essential to the interpretation of the data but inaccessible to present methods. The standard maximum likelihood methods for analyses were developed and tested on COBE measurements (e.g., G\\'orski 1994, G\\'orski \\etal 1994, 1996, Bond 1995, Tegmark \\& Bunn 1995, Hinshaw \\etal 1996, Tegmark 1996, Bunn \\& White 1996, Bond, Jaffe, \\& Knox 1998, hereafter BJK98, 2000) which have only about $N \\simeq 1000$ pixels. Balloon-borne experiments, such as Boomerang (de Bernardis \\etal 2000), Maxima (Hanany \\etal 2000), and Tophat (Martin \\etal 1996), and ground based measurements e.g., TOCO (Miller \\etal 1999) and Viper (Peterson \\etal 2000) with up to $N \\simeq 10^5$ are already pushing present day supercomputer technology. The future missions MAP and Planck with $N = 10^6-10^7$ are estimated to require up to millions of years (Borrill 1999, Bond, Crittenden, Jaffe, \\& Knox 2000) for one iteration of the quadratic maximum likelihood estimator for $C_l$'s. The disk storage and memory usage of these algorithms are prohibitive as well. To date fast algorithms were presented under two fairly restrictive assumptions: i) the noise is both temporally uncorrelated and spatially axially symmetric ii) the foregrounds can be exactly removed both from the map and the correlation matrix (Oh, Spergel, \\& Hinshaw 1999, Wandelt, Hivon \\& Gorski 1998, 2000). Our approach is significantly different from these, since it does not assume any symmetries. ", "conclusions": "We presented a novel method to extract ${\\cal C}_\\ell$'s from large CMB maps via two-point correlation functions. We have cast our determination of ${\\cal C}_\\ell$ into 3 steps: fine-grained $C(\\theta)$ estimation, the pixel-related $4^\\prime$ Gaussian smoothing of it, followed by $P_\\ell$ multiplication and Gauss-Legendre integration. The resulting expression for ${\\cal C}_\\ell$ can be itself understood as a sum over pixel pairs, with a pair-weighting proportional to the Gaussian smoothing of $P_\\ell$. In this respect our estimator belongs to the class of quadratic estimators for ${\\cal C}_\\ell$, of which other examples (e.g., BJK98) have been used in the past and are expected to give similar results. In the present example the measured errorbars were at most $10$\\% larger then the theoretically minimal ones; this negligible suboptimality allowed the reduction of the CPU time from months (Borrill 1999, Table 3) to hours. For future large missions the difference will be even more dramatic: optimal methods would improve the errorbars only slightly at a prohibitive cost. Our new approach has several advantages in comparison to previous techniques. It scales as $N^2$ even in its most straightforward implementation, compared to typical quadratic estimators which scale as $N^3$. This scaling can be further improved up to $N\\log N$. While in the present implementation we used diagonal noise, no other (e.g., asymuthal) symmetries were assumed about noise, or geometry of the map. Cut out holes around bright sources, galactic cut, any irregularity in the sampling, make no difference in the speed or performance of the algorithm. To illustrate this, Equation~\\ref{eq:estimator} was used to determine the correlation function from COBE DMR data, which has inhomogeneous noise. The recovered $C_l$'s are consistent with those obtained from implementations of quadratic estimators discussed in BJK98. In contrast with any other previous attempt, our approach is straightforward to generalize for non-diagonal noise matrix (see below). While our method in its present form is already practical for analyzing megapixel CMB maps, it has the potential for further generalizations. General non-diagonal noise appears to pose a serious problem. Fast iterative map making methods (Wright \\etal 1996, Prunet \\etal 2000) capable of handling large data sets only furnish the weight matrix, $w_{ij} = N_{ij}^{-1}$. Since inversion of the noise matrix is again an $N^3$ problem, our estimator of Equation~\\ref{eq:estimator} might not be directly applicable. Instead, we propose a new estimator using MC realizations of artificial noise \\begin{equation} \\tilde\\xi(\\cos\\theta) = \\sum_{ij} f_{ij} (\\Delta_i \\Delta_j - \\frac{1}{M}\\sum_{k=1}^M n_i^k n_j^k), \\label{eq:modestimator} \\end{equation} where $n_i^k$ is one of $M$ realizations of the noise for pixel $k$. This plays the role of a MC inversion of the weight matrix $w_{ij} = N_{ij}^{-1}$. Generation of $n_i^k$ in the time domain, where it has simple correlation structure, is straightforward. Its iterative projection into pixel space is equivalent to map-making. This is feasible even when storing the noise-matrix would be prohibitive, as for Planck. To further improve the performance of our method, we will implement a minimum variance weighting scheme (Feldman Kaiser, \\& Peacock 1994, and Colombi, Szapudi, \\& Szalay 1998). This corresponds to down-weighting measurements with their variance, and might be important in maps, where various pairs contributing to the correlation function have widely differing errors. Otherwise, the uniform weighting scheme is nearly optimal (Colombi, Szapudi, \\& Szalay 1998). The present simple weighting scheme can be improved heuristically via individual pixel weighting reflecting differences in sky coverage. More complex pair weighting can be defined iteratively, using the MC estimates of the variances. The present Letter used a simple $N^2$ code to calculate the correlation function, and this is perfectly adequate for LDB surveys with $N\\simeq 10^5$ pixels, and, with supercomputers, even for MAP. Nevertheless, the development of an $N\\log N$ code is under way (Szapudi \\& Colombi 2000, Connolly, Nichol, Moore, \\& Szapudi 2000). Our technique has reestablished the utility of correlation functions for CMB studies. This approach has further potential applications: it is naturally generalizable for the assessment of non-Gaussianity in CMB maps via $k$-point correlation functions, with implementations as fast as $N(\\log N)^{k-1}$ (Sunyaev-Zeldovich effect, lensing of CMB); it is equally useful for polarization correlation functions and for obtaining the appropriate $C_l$'s from them. We have found that the statistical information is condensed into singular features of the correlation function, (see also Bashlinsky \\& Bertschinger, in prep.). This suggests that direct parameter estimation from the CMB two-point correlation function might be fruitful as well. Other applications include correlations of the infrared (SCUBA, BLAST, SIRTIF, FIRST) and optical background, and weak gravitational lensing." }, "0010/astro-ph0010604.txt": { "abstract": "We use a multi-color classification method introduced by Wolf, Meisenheimer \\& R\\\"oser (2000) to reliably identify stars, galaxies and quasars in the up to 16-dimensional color space provided by the filter set of the Calar Alto Deep Imaging Survey (CADIS). The samples of stars, galaxies and quasars obtained this way have been used for dedicated studies which are published in separate papers. The classification is good enough to detect quasars rather completely and efficiently without confirmative spectroscopy. The multi-color redshifts are accurate enough for most statistical applications, e.g. evolutionary studies of the galaxy luminosity function. Also, the separation between stars and galaxies reaches deeper than with morphological criteria, so that studies of the stellar population can be extended to fainter levels. We characterize the dataset presently available on the CADIS 1h-, 9h- and 16h-fields. Using Monte-Carlo simulations we model the classification performance expected for CADIS. We present a summary of the classification results on the CADIS database and discuss unclassified objects. More than 99\\% of the whole catalog sample at $R<22$ (more than 95\\% at $R<23$) are successfully classified matching the expectations derived from the simulations. A small number of peculiar objects challenging the classification are discussed in detail. Spectroscopic observations are used to check the reliability of the multi-color classification (6 mistakes among 151 objects with $R<24$). From these, we also determine the accuracy of the multi-color redshifts which are rather good for galaxies ($\\sigma_z \\approx 0.03$) and useful for quasars. We find that the classification performance derived from the simulations compares well with results from the real survey. Finally, we locate areas for potential improvement of the classification. %Black \\& white seems to be out, now that color vision is established in astronomical households. More and more stars prefer to be featured in color presentations, dressing in all thermal colors nowadays. Even black becomes fashionable, being still colorful in innovative near infrared. A brief explanation of modern color trends follows. % ", "introduction": "The {\\it Calar Alto Deep Imaging Survey} (CADIS) is an extragalactic key project at the Max-Planck Institut f\\\"ur Astronomie (MPIA), Heidelberg, which is aiming at two types of objectives: CADIS investigates whole samples of different object classes using statistical tools, but it also searches for individual rare and faint objects, which will be studied in detail with coming large telescopes. As a pencil beam survey, it probes seven different fields at galactic latitudes $b \\ga 45\\degr$ with a total area of $\\sim 0.25\\degr$. The final object catalog will arise from two fundamentally different survey techniques: \\begin{itemize} \\item a multi-color survey with B, R, J and K$^\\prime$ plus 13 medium-band filters from 400\\,nm to 1000\\,nm, practically resembling low-resolution imaging spectroscopy and giving a complete list of objects with $R \\la 23$, \\item and an emission-line survey using an imaging Fabry-Perot interferometer to probe emission line galaxies down to a limiting line flux of $\\sim 3 \\times 10^{-20} Wm^{-2}$. \\end{itemize} Presently, the data for three fields are reduced and have been used for a number of application studies published already or to be published this year. For many applications, objects of concern are selected by our multi-color classification which uses the many bands to sort the objects into {\\it stars}, {\\it galaxies} and {\\it quasars}. Also, multi-color redshifts are estimated for the extragalactic objects. This classification scheme was originally developed on the basis of CADIS data, but is meanwhile used in a range of different survey activities. It uses a library of $\\sim 65000$ templates and achieves high classification reliability ($>90\\%$ correct classification in each class) and a high redshift accuracy of $\\sigma_z \\approx 0.03$ for galaxies and $\\sigma_z \\approx 0.1$ for quasars. The methodogical background for the classification was published by Wolf, Meisenheimer \\& R\\\"oser (2000), hereafter paper I. There, the classification method is derived from statistical principles, the libraries are defined, the performance expected with different filter sets is compared and where conclusions are drawn for optimum survey strategies. The purpose of this paper is to characterize the present data of the CADIS multi-color survey and to discuss the classification performance which was checked by a subsample of objects with spectroscopic identifications. The paper is organized as follows: Section 2 lists CADIS goals for which the classification is relevant and discusses what kind of objects the classifications should be prepared for. Section 3 defines the present CADIS dataset and characterizes its photometry and calibration. Section 4 outlines the classification method and presents statistics of classified and unclassified objects. Section 5 discusses the classification performance using a spectroscopic cross check sample. Finally, Section 6 summarizes the quality of the classified catalogs and evaluates the practical performance of the CADIS multi-color classification. A few identified peculiar objects challenging the classification are discussed in the appendix. ", "conclusions": "In paper I an innovative method for identifying stars, galaxies and quasars in multi-color surveys was presented, which uses a library of $\\ga 65000$ color templates for comparison with observed objects. The method aims for extracting the information content of object colors in a statistically correct way, and performs a classification as well as a redshift estimation for galaxies and quasars in a unified approach based on the same probability density functions. The three basic ingredients to this method are \\begin{enumerate} \\item accurately measured and calibrated color data for the objects to be classified including the color errors \\item accurate and representative color template libraries covering the range of objects expected in the dataset obtained \\item a statistical classifier and parameter estimator which can potentially be trimmed for best performance in particular applications. \\end{enumerate} Also, in paper I it was concluded that medium-band surveys are expected to deliver a performance superior to pure broad-band surveys even under the constraint of equally limited telescope time. Based on survey simulations this method should be capable of \\begin{enumerate} \\item separating stars from galaxies down to deeper limits than possible by using morphology only \\item finding quasars also at redshifts where its colors overlap with the stellar locus in two-color projections, particularly in the range of $2.22$ was confirmed by spectroscopic observations \\cite{Wolf99}. The fraction of unclassified objects is less than 1\\% at $R<22$ and reaches about 50\\% at $R=24$ due to increasing photon noise that tends to make different original SEDs equally likely sources of the observed colors. At some level in between the classification becomes incomplete and also features an increasing fraction of actual mistakes. These are basically all members of the rich galaxy population spilling over into regions of color space that is usually occupied by stars and quasars. A spectroscopic cross check using 162 identifications confirmed the multi-color classification to work essentially free of errors at $R<22$ (two mistakes among 103 objects). We do not have proper knowledge about where and how the classification collapses exactly, which is particularly important for the rare quasar class which will become incomplete and dominated by contaminating galaxies at some level. To settle this uncertainty, dedicated spectroscopic observations are required. These would be a valuable and not too time-consuming investment, if a large number of spectra from a CADIS size field could be taken simultaneously with an instrument like VMOS at the VLT. The findings on the classification performance are consistent with the expectations raised by the Monte-Carlo simulations of the CADIS filter set. They suggest the classification to work nearly perfect down to $R \\approx 23$ except for some contamination of quasar candidates by emission line galaxies towards that limit. The simulations actually appear to be too optimistic about the working depth of the classification by about a third of a magnitude. Of course, a real survey will perform worse than a simulation, that can not account for differences between the real world and our library and does not contain artifacts and variable or blended objects. The latter issues have to be dealt with better data analysis. For galaxies, our multi-color redshifts are useful down to $R \\sim 24$. The statistics on the redshift errors are dominated by $\\sim 10\\%$ catastrophic mistakes, where the estimator decides for the wrong one among alternative values with comparable probability. The core of the error distribution has a zero mean error and $\\sigma_z \\approx 0.03$ rms width. Quiescent galaxies tend to work better than starburst types. Half of the quasars receive remarkably correct estimates with an average variance of $\\sigma_z \\approx 0.1$, preferentially the $z>2$-objects and those of higher luminosity. In contrast, a large amount of redshift confusion is expected at lower redshift. Also, simulations for quasars work better than the real redshift estimation, probably mostly due to variability. Eventually, the classified multi-color catalog has been used for several dedicated studies published in separate papers: \\begin{enumerate} \\item The star-galaxy separation with the multi-color data reaches deeper than a morphological separation. Phleps et al. (2000) could therefore use the sample of stars to probe the stellar content and the Galactic structure along the pencil beams established by the CADIS fields, where they find strong evidence of a thick disk. \\item The multi-color redshifts of galaxies are sufficiently accurate for most statistical studies. Also, the classification includes a substantial fraction of compact galaxies into the sample. Using that, Fried et al. (2000) investigated the evolution of galaxies within $0.32$-quasars. Many of these might reside close to the stellar locus and be overlooked in broad-band surveys \\cite{Wolf99}. Thus, quasars can be identified with a much more uniform completeness across the accessible redshift range and more homogeneous samples can be obtained. At $z>2$, it will be possible to constrain the evolution of the luminosity function from a large enough multi-color sample. \\end{enumerate}" }, "0010/hep-ph0010280_arXiv.txt": { "abstract": "The initial condition problem of inflation is examined from the perspective of both spacetime embedding and scalar field dynamics. The spacetime embedding problem is solved for arbitrary initial spatial curvature $\\Omega$, which generalizes previous works that primarily treat the flat case $\\Omega=1$. Scalar field dynamics that is consistent with the embedding constraints are examined, with the additional treatment of damping effects. The effects of inhomogeneities on the embedding problem also are considered. A category of initial conditions are identified that are not acausal and can develop into an inflationary regime. ", "introduction": "\\label{introduction} Inflation is the foremost idea for explaining the large scale homogeneity and near flatness of the universe, which are the two main unexplained observational features in the standard hot big-bang model. The large scale homogeneity or horizon problem amounts to the fact that under hot big-bang evolution, due to a decelerating scale factor ${\\ddot a} < 0$, sufficiently separated regions of the present day observable universe would never have been in causal contact. Nevertheless, it appears the universe we observe looks very much the same, and in particular very smooth, in all directions. This fact is best seen in the cosmic microwave background radiation (CMB) which has temperature fluctuations of only one part in $10^5$ when measured from any direction in the sky \\cite{Bennett:1996ce}. The inflation solution to this horizon problem is to picture the universe during an early epoch to undergo an accelerated expansion, ${\\ddot a} > 0$. Such expansion can take an initially small causally connected patch and enlarge it to a size that comfortably encompasses our present day observed universe, thereby solving the horizon problem. To realize inflation, the equation of state within the inflationary patch must be of a very special form, possessing negative pressure. The potential energy of a scalar field has an equation of state that satisfies this requirement. This fact has been a key link towards a dynamical realization of inflation and thereby has further motivated the inflation solution. As inflation is meant to solve the horizon problem, it is important that it does not require acausal initial conditions. In particular, the picture of inflation considered here is for the universe to emerge from an initial singularity and then enter into a hot big-bang radiation dominated evolution. At some time $t_i$ after the initial singularity, the conditions appropriate for inflation should occur within a small patch that is contained within the causal horizon at that time. Chaotic inflation \\cite{ci,Linde:1990nc} does not fall into this picture as there inflation is thought to start at the Planck epoch with homogeneity assumed on the Planck scale. To realize the picture of a ``local'' inflation, two requirements must be satisfied. {}First a physically sensible embedding must be demonstrated of the inflating patch immersed within a non-inflating background \\cite{embedd,Vachaspati:2000dy,Trodden:1999wc}. By embedding we mean matching the inflationary space time with the background space time at the boundary of the patch. Second, it must be shown that for $t> t_i$, the patch is dynamically stable to sustain inflation \\cite{piran181,kb,kb90,gpprl,gold43,other,Goldwirth:1992rj}. {}For example, large fluctuations, which conceivably could enter the inflating patch at a maximum rate limited only by causality, should not destroy the inflationary conditions within the patch. Recently, a convenient methodology has been developed in \\cite{Vachaspati:2000dy,Trodden:1999wc} for analyzing the embedding problem, based on the null Raychaudhuri equation \\cite{ray,Wald:1984rg}. In Sec. \\ref{embedding_conditions} the flat spacetime ($\\Omega=1$) formulation in \\cite{Vachaspati:2000dy} is generalized to arbitrary spacetime curvature. Then, an alternative solution from \\cite{Vachaspati:2000dy} to the embedding problem will be identified, which is especially attractive for an inflating patch with an open geometry. In Sec. \\ref{dynamic_conditions}, initial conditions for scalar field dynamics are presented, which are consistent with causality and our embedding solution and which evolve into successful supercooled or warm inflationary regimes. In Sec. \\ref{dynamical_effects_on_embedding} we combine the results in Sec. \\ref{embedding_conditions} and Sec. \\ref{dynamic_conditions} to evaluate the effect of inhomogeneities on the embedding problem. Finally, Sec. \\ref{conclusion} presents our conclusions. ", "conclusions": "\\label{conclusion} This paper has investigated the initial condition problem of inflation from the perspective both of spacetime embedding and inflaton dynamics. Our study has highlighted two attributes of this problem which have not been addressed in other works. {}First, from the perspective of spacetime embedding, we have observed that the global geometry can play an important role in determining the size of the initial inflationary patch that is consistent with the weak energy condition. Second, from the perspective of inflaton dynamics, we have noted that a $\\Gamma {\\dot \\phi}$ damping term could alleviate several problems which traditionally have led to large scale homogeneity requirements before inflation. The purpose of this paper was to note for both these attributes, their salient features with respect to the initial condition problem. In the wake of this, several details emerge that must be understood. Below, we will review the main result we found for both attributes and then discuss the questions that must be addressed in future work. {}For a causally generated patch a successful embedding can be achieved if the patch does not contain an anti-trapped region. We have shown that the MAS size can be arbitrarily larger than the Hubble scale provided $\\Omega$ is made small enough. So if the patch Hubble horizon is taken as the minimum stable patch size, then the patch does not have to contain an anti-trapped region if $\\Omegainf < 1$. This generalizes the analysis of \\cite{Vachaspati:2000dy} which was only for $\\Omega=1$. However, without the effects of damping, it appears that the event horizon, not the Hubble horizon, is the minimum stable patch size. {}For the de Sitter case one can see analytically that the event horizon is equal to the MAS size regardless of $\\Omega$ and numerical calculations suggest the same is true for power law inflation. However, radiation damping of perturbations could stabilize a patch smaller than the event horizon. In this case an open geometry for the patch would allow the patch not to contain an anti-trapped region and thus allow a causal embedding in an expanding background which does not violate the weak energy condition. Eventually the patch should develop a MAS within it, but by then it could have expanded to be larger than the background MAS. {}For the dynamics problem, with respect to the scalar field the new consideration was the effect of a $\\Gamma {\\dot \\phi}$ damping term. We found that such a term could suppress many of the effects from initial inhomogeneities of the inflaton, which in studies traditionally done without this term lead to important impediments to entering the inflation regime. It appears evident that inclusion of such a damping term will lead to qualitative differences in the initial condition problem. The most interesting outcome is initial inflationary patches smaller than the Hubble radius $1/{\\rm H}$ may be able to inflate. This paper examined the consequences of damping terms but did not delve into their fundamental origin. {}For the cosmological setting, such damping terms are typically associated with systems involving a scalar field interacting with fields of a radiation bath. In this case, such damping terms have been found in first principles calculations for certain warm inflation models \\cite{wifp}, although more work is needed along these lines. It is worth noting here that in the early stages of certain supercooled inflaton scenarios where radiation is present, in particular new \\cite{ni} and thermal \\cite{ti} inflation, a careful examination of the dynamics may reveal damping terms similar to this. Since the initial stages are the crucial period for the initial condition problem, if further study supports the importance of such damping terms, it may be useful to better understand damping effects also in such scenarios. Since the most suggestive situation for the damping terms is where in addition a radiation component is present in the universe, in Subsect. IIIB we also studied the effects of this component on the initial condition problem. Specifically we studied the case most suggestive for the initial stages of new and warm inflation, where a small inflation patch is submerged inside a larger radiation dominated spatial region. Our main observation has been that the minimal condition for the patch to inflate is that its vacuum energy density must be larger than the background radiation energy density. Provided the vacuum energy sustains itself, since expansion of the universe will dilute the radiation energy density, it is not clear-cut that the external radiation energy can act with sufficient magnitude to impede inflation inside the patch. We have shown that the gradient terms, due to the inhomogeneous scalar field in the background space time, make it possible for the patch boundary to overtake the background MAS. However the kinetic energy of the scalar field must not be dominant for this to happen. This can be ensured by including a damping term in the scalar field equation." }, "0010/astro-ph0010382_arXiv.txt": { "abstract": "Spectroscopic observations of a proto-planetary nebula He 3-1475 with the Space Telescope Imaging Spectrograph (STIS) reveal the kinematics of its high (1200 km s$^{-1}$) velocity jets. The jets are formed at a large (0.15 pc) distance from its central star by collimation of an asymmetric stellar wind in a pair of conical shocks seen in Wide Field Planetary Camera (WFPC2) images. The jets consist of several pairs of knots symmetrically distributed with respect to the central star, with most knots exhibiting a head-tail morphology. Large (up to 650 km s$^{-1}$) radial velocity gradients are seen within the knots on subarcsec scales, with velocities decreasing from the knot heads toward their trailing tails. These large velocity gradients are a sign of efficient deceleration of jets by a much slower bipolar outflow. The inclination angle of the bipolar outflow is equal to 40\\degr, based on Doppler shifts of the scattered stellar H$\\alpha$ line. Its velocity is equal to 140 km s$^{-1}$ at a distance of 0.23 pc from the star, and increases monotonically with the radial distance from the star. A comparison of new WFPC2 [N II] $\\lambda$6584 images with older WFPC2 images reveals expansion of the jets. The measured jet proper motions in combination with their radial velocities imply that He 3-1475 is a Galactic Bulge star at a distance of 8 kpc, located 800 pc above the Galactic plane. Its very high luminosity (25,000 $L_\\odot$) implies that He 3-1475 must be significantly more massive than a typical AGB star within the Galactic Bulge, perhaps because of a past mass transfer and/or a merger event in an interacting binary system. ", "introduction": "He 3-1475 is a B[e] type star located in the direction of the Galactic Bulge, at $l = 9$\\fdg4, $b = +5$\\fdg8. Stars of this spectral type do not show typical absorption line spectra, but instead numerous permitted and forbidden emission lines are superposed on featureless continua (for further information see a recent review on B[e] stars by Zickgraf 1998). He 3-1475 is an unusual B[e] star because it is surrounded by a dense expanding torus of circumstellar matter (CSM) seen in OH maser emission (Bobrowsky et al. 1995), and also as a prominent dark torus in Hubble Space Telescope (HST) images (Bobrowsky et al. 1995; Borkowski et al. 1997). A bipolar outflow and spectacular jets, perpendicular to the torus, were detected in the optical by Riera et al. (1995) and Bobrowsky et al. (1995). Radial velocities of bright knots in jets span 1000 km s$^{-1}$, which is an order of magnitude more than what is typically seen in both young and evolved stars surrounded by dense CSM. The morphology of the jets is also unusual, because they do not originate in the central star. Instead, a fast bipolar stellar outflow is apparently collimated far from the star in conical shocks seen in the HST images (Borkowski et al. 1997). The jets above the tips of conical shocks consist of a series of bright pairs of knots, located symmetrically with respect to the central star. Their spectra show the presence of substantial ($\\sim 450$ km s$^{-1}$) velocity gradients within the knots, which indicates their strong deceleration by the ambient, more slowly moving material in the bipolar lobes (Riera et al. 1995). There is also evidence for deceleration of knots on large scales, as the knot radial velocities decrease with increasing distance from the cone tips to the periphery of the bipolar outflow. He 3-1475 is clearly an unusual source whose nature is not understood at present. It's distance has been estimated at 1 -- 5 kpc (Riera et al. 1995; Bobrowsky et al. 1995), which together with its high infrared flux (Parthasarathy \\&\\ Pottasch 1989) suggests that He 3-1475 is probably a post-Asymptotic Giant Branch (AGB) proto-planetary nebula (PPN). (Knapp et al. 1995 detected 8.4 GHz continuum radio emission, which they interpreted as coming from a compact, newly formed PN close to the central star.) This conclusion should be considered preliminary because a reliable distance determination is lacking. The geometry of the circumstellar medium is also poorly known, with an uncertain (30\\degr\\ -- Riera et al. 1995) inclination of the jets with respect to the line of sight. The velocity of the bipolar outflow and its age are unknown. This lack of fundamental knowledge about He 3-1475 makes it difficult to understand this puzzling object. We report here on imaging and spectroscopic observations of He 3-1475 with the HST, which allow us to determine its distance, inclination and kinematics of its jets, and the velocity of the bipolar outflow. Imaging and astrometry with the Wide Field Planetary Camera (WFPC2) and the Space Telescope Imaging Spectrograph (STIS) data are presented and discussed in \\S~2 and \\S~3, respectively. The jet inclination and the distance to He 3-1475 are determined in \\S~4 and \\S~5, and the nature of He 3-1475 and of its jets are discussed in \\S~6. ", "conclusions": "Our distance determination places He 3-1475 in the Galactic Bulge, at a much larger distance than previously anticipated. At this distance its infrared luminosity is equal to 25,000 $L_\\odot$ (Riera et al. 1995), which might be less that the stellar luminosity if the dust optical depth is low along the bipolar outflow so that a substantial fraction of stellar luminosity escapes along the poles. He 3-1475 is apparently a high-luminosity Galactic Bulge star at a height of 800 pc above the Galactic plane. Its location suggests that He 3-1475 is an old Population II star, but an old, low-mass post-AGB star cannot be so luminous -- a more massive star is required. Luminous, more massive AGB stars are indeed found in old stellar populations (e. g., in the globular cluster NGC 6553 -- Guarnieri et al. 1997), although their luminosities may be several times lower than that of He 3-1475. It is also feasible that He 3-1475 is a member of a trace intermediate-age population of Galactic Bulge. Irrespective of its age and origin, He 3-1475 appears to be an evolved star which lost a substantial amount of mass in a slow asymmetric wind at least 1,500 yr ago, possibly in a superwind at the AGB tip or during a binary interaction, and which is now seen in the optical as a dense torus. The presence of broad P-Cygni profiles in the stellar spectrum shows that the transition from the slow to the fast wind has occurred since then. The B[e] spectral classification of the central star implies that this fast wind is asymmetric, according to a disk wind model for B[e] stars by Zickgraf et al. (1985). This wind is perhaps driven by stellar UV radiation from a single fast rotating B type star (e. g., Pelupessy, Lamers, \\& Vink 2000), or more generally from a star + a disk system (Oudmaijer et al. 1998). This asymmetric wind is then collimated into 1200 km s$^{-1}$ jets during its interaction with the more slowly moving material ejected earlier, maybe purely by hydrodynamical means (Borkowski et al. 1997). The time variability often seen in many B[e] stars suggests that a ``knotty'' jet morphology seen in He 3-1475 could be caused by temporal variations in the stellar wind properties. Our STIS observations show how this high-velocity, time-dependent jet is decelerated by the circumstellar medium on both large and small scales. We see the large scale deceleration in the decrease of the maximum velocity as we move down the jet and out to the knots. But we also see at all points along the flow, and especially at the knots, local ($\\sim 0\\farcs05 = 6 \\times 10^{15}$ cm) decelerations of over 500 km s$^{-1}$. Consider that a plane-parallel shock of 500 km s$^{-1}$ would produce gas with a temperature of $3 \\times 10^6$ K, while we see a spectrum which seems more appropriate for a cooling shock of far lower velocity, about 100 km s$^{-1}$ (Riera et al. 1995). Clearly, there is much hydrodynamical structure which is still unresolved by HST. It is still a mystery why B[e] star+disc systems are present in evolved post-AGB stars such as He 3-1475. There are a dozen such systems in compact PNe or proto-PNe (Lamers et al. 1998), albeit generally of lower luminosity than He 3-1475. The best known of these systems is the ``Butterfly'' PN M2-9, a low luminosity (550 L$_\\odot$) object at a distance of 650 pc (Schwarz et al. 1997). M2-9 is clearly very much different than He 3-1475, suggesting that compact PN B[e] central stars form a heterogeneous class, just like B[e] stars as a whole (Lamers et al. 1998). M2-9 shows a complex kinematics, with young ($< 10$ yr) micro-jets close to the star with velocities up to 195 km s$^{-1}$ and much older (1,300 yr) bipolar outflows with velocities up to 140 km s$^{-1}$ (Solf 2000). The 1200 yr old outermost reflection knots are expanding with velocity of 165 km s$^{-1}$ (Schwarz et al. 1997). Doyle et al. (2000) find evidence for $\\sim 1000$ km s$^{-1}$ outflows. High velocity (500 km s$^{-1}$) outflows were also found by Redman et al. (2000) in another PN with a B[e] central star, Mz3, while spectacular jets were recently imaged by the HST in yet another PN on Lamers' et al. PN and PPN list, He 2-90 (Sahai \\& Nyman 2000). Both M2-9 and He 2-90 are suspected to be interacting binaries. An interacting binary progenitor hypothesis also seems attractive for He 3-1475, because its very high luminosity excludes a typical low-mass AGB star in the Galactic Bulge. A mass transfer binary and/or a merger event provides a reasonable solution to the presence of a more massive star in the Galactic Bulge. In this paper, we concentrated our attention on the large-scale jet kinematics and on the global properties of He 3-1475. The next step in analysis of the extensive STIS dataset should involve detailed studies of individual knots and of the jet collimation regions. Such studies are essential for providing answers to questions about the mechanism of the jet collimation, about the nature of the observed shock emission, and about the temporal and spatial variability of the flow. For example, a simple (but perhaps naive) interpretation would be to attribute the observed shock emission to shocks driven into the jets, such as encountered in working surfaces of continuous jets. Because of the high jet velocities, the bow shocks ahead of the knots should in principle be nonradiative, X-ray emitting shocks, with little or no optical emission. The presence of distinct knots symmetrically distributed with respect to the star suggests that the stellar outflow varies on timescales from tens to hundreds of years, leading to noncontinuous, episodic, and wobbling jets. It is also not certain whether the jets can be collimated by purely hydrodynamical processes, or whether magnetic fields must be involved. We hope that the continuing analysis of the HST He 3-1475 dataset will provide answers to some of these questions in the near future." }, "0010/astro-ph0010457_arXiv.txt": { "abstract": "We model the population of double white dwarfs in the Galaxy and find a better agreement with observations compared to earlier studies, due to two modifications. The first is the treatment of the first phase of unstable mass transfer and the second the modelling of the cooling of the white dwarfs. A satisfactory agreement with observations of the local sample of white dwarfs is achieved if we assume that the initial binary fraction is $\\sim$50 \\% and that the lowest mass white dwarfs ($M < 0.3 \\msun$) cool faster than the most recently published cooling models predict. With this model we find a Galactic birth rate of close double white dwarfs of 0.05 yr$^{-1}$, a birth rate of AM CVn systems of 0.005 yr$^{-1}$, a merger rate of pairs with a combined mass exceeding the Chandrasekhar limit (which may be progenitors of SNe~Ia) of 0.003 yr$^{-1}$ and a formation rate of planetary nebulae of 1 yr$^{-1}$. We estimate the total number of double white dwarfs in the Galaxy as 2.5\\,$\\times\\,10^8$. In an observable sample with a limiting magnitude $V_{\\rm lim} = 15$ we predict the presence of $\\sim$855 white dwarfs of which $\\sim$220 are close pairs. Of these 10 are double CO white dwarfs of which one has a combined mass exceeding the Chandrasekhar limit and will merge within a Hubble time. ", "introduction": "Close double white dwarfs\\footnote{Throughout this work we'll use the term double white dwarf instead of double degenerate, which is commonly used, because the term double degenerate is sometimes used for white dwarf - neutron star or double neutron star binaries.} form an interesting population for a number of reasons. First they are binaries that have experienced at least two phases of mass transfer and thus provide good tests for theories of binary evolution. Second it has been argued that type Ia supernovae arise from merging double CO white dwarfs \\citep{web84,it84a}. Thirdly close double white dwarfs may be the most important contributors to the gravitational wave signal at low frequencies, probably even producing an unresolved noise burying many underlying signals \\citep{eis87,hbw90}. A fourth reason to study the population of double white dwarfs is that in combination with binary evolution theories, the recently developed detailed cooling models for (low-mass) white dwarfs can be tested. The formation of the population of double white dwarfs has been studied analytically by \\citet{it86a,it87} and numerically by \\citet{lp88,ty93,ty94,ylt+94,hpe95}; \\citet[hereafter \\citetalias{ity97}]{ity97}, and \\citet[hereafter \\citetalias{han98}]{han98}. Comparison between these sudies gives insight in the differences that exist between the assumptions made in different synthesis calculations. Following the discovery of the first close double white dwarf \\citep{slo88}, the observed sample of such systems in which the mass of at least one component is measured has increased to 14 \\citep{mm99,mmm+00}. This makes it possible to compare the models to the observations in more detail. In this paper we present a new population synthesis for double white dwarfs, which is different from previous studies in three aspects. The first are some differences in the modelling of the binary evolution, in particular the description of a common envelope without spiral-in, in which the change in orbit is governed by conservation of angular momentum, rather than of energy (Sect.~\\ref{binev}). The second new aspect is the use of detailed models for the cooling of white dwarfs (Sect.~\\ref{cooling}), which are important because it is the rate of cooling which to a large extent determines how long a white dwarf remains detectable in a magnitude-limited observed sample. The third new aspect is that we use different models of the star formation history (Sect.~\\ref{SFH}). Results are presented in Sect.~\\ref{results} and discussed in Sect. \\ref{discussion}. The conclusions are summarised in Sect.~\\ref{conclusion}. In the Appendix some details of our population synthesis are described . ", "conclusions": "We computed a model of the population of close binary white dwarfs and found good agreement between our model and the observed double white dwarf sample. A better agreement with observations compared to earlier studies is found due to two modifications. The first is a different treatment of unstable mass transfer from a giant to a main sequence star of comparable mass. The second is a more detailed modelling of the cooling of low mass white dwarfs which became possible because detailed evolutionary models for such white dwarfs became available. Our main conclusions can be summarised as follows. \\begin{enumerate} \\item Comparing the mass distribution of the white dwarfs in close pairs with the observations, we find a lack of observed white dwarfs with masses below 0.3 \\msun. This discrepancy can be removed with the assumption that low-mass white dwarfs cool faster than computed by \\citet{dsb+98}. The same assumption removes discrepancies between observed and derived ages of low-mass white dwarfs that accompany recycled pulsars, as shown by \\citet{kbk+00}. Faster cooling is expected if the hydrogen envelopes around low-mass white dwarfs are partially expelled by thermal flashes or a stellar wind. \\item Our models predict that the distribution of mass ratios of double white dwarfs, when corrected for observational selection effects as described by \\citet{mmm00}, peaks at a mass ratio of unity, consistent with observations. The distributions predicted in the models by \\citet{ity97} and \\citet{han98} peak at mass ratios of about 0.7 and above 1.5 and agree worse with the observations even after applying selection effects. \\item Our models predict a distribution of orbital periods and masses of close double white dwarfs in satisfactory agreement with the observed distribution. \\item Amongst the observed white dwarfs only a small fraction are members of a close pair. To bring our models into agreement with this, we have to assume an initial binary fraction of 50\\%\\ (i.e.\\ as many single stars as binaries). \\item In our models the ratio of the local number density of white dwarfs and the planetary nebula formation rate is a sensitive function of the star formation history of the Galaxy. Our predicted numbers are consistent with the observations. \\item Using detailed cooling models we pridict that an observed sample of white dwarfs near the Sun, limited at the magnitude $V=15$, contains 855 white dwarfs of which 220 are close pairs. Of these pairs only 10 are double CO white dwarfs and only one is expected to merge having a combined mass above the Chandrasekhar mass. The predicted merger rate in the Galaxy of double white dwarfs with a mass that exceeds the Chandrasekhar mass is consistent with the inferred SN~Ia rate. \\citetalias{ity97} estimated, depending on $\\alpha_{\\rm ce}$, to find one such pair in a sample of $\\sim$200 to $\\sim$600 white dwarfs. Reversing this argument, when the statistics become more reliable, the observed number of systems with different types of white dwarfs could provide constraints on the cooling models for these white dwarfs. \\end{enumerate}" }, "0010/astro-ph0010331_arXiv.txt": { "abstract": "To build upon the goals of the upcoming INTEGRAL mission, the next generation soft $\\gamma$-ray (0.2-20~MeV) observatory will require improved sensitivity to nuclear line emission while maintaining high spectral resolution. We present the simulated performance of a germanium Compton telescope (GCT) design, which will allow a factor of ten improvement in sensitivity over INTEGRAL/SPI. We also discuss a number of issues concerning reconstruction techniques and event cuts, and demonstrate how these affect the overall performance of the telescope. ", "introduction": "Combining the success of COMPTEL/CGRO and the spectral resolution of INTEGRAL/SPI, a number of researchers \\citep{johns96, jean96, boggs98} have discussed the merits of a germanium Compton telescope (GCT). Compton telescopes work on a well-known principle: by measuring the positions and energies of the photon interactions the initial photon direction can be reconstructed to within an annulus on the sky using the Compton scatter formula (Figure 1). The uncertainty, or width, of this annulus depends on the spatial and spectral resolution of the detectors, but also has a fundamental limit set by Doppler broadening due to Compton scattering off of bound electrons. Reconstruction of the event annulus requires that the first and second photon interaction locations in the instrument are spatially resolved, and their order properly determined. Germanium detectors pose two major complications for Compton telescope designs. Photons above $\\sim$0.5~MeV predominantly scatter multiple times in germanium before being photoabsorbed. Also, the expected event timing resolution in germanium detectors ($>$10~ns) is not adequate to determine the interaction order for reasonable GCT configurations. \\begin{figure} \\centering \\includegraphics[width=1.0\\linewidth]{figure1.eps} \\caption{Germanium Compton telescope configuration analyzed in this work.\\label{figure1}} \\end{figure} In a previous paper \\citep{boggs00}, hereafter Paper I, we introduced two \\textit{reconstruction techniques} to accurately determine the photon interaction order in GCTs. The first technique, Compton Kinematic Discrimination (CKD), takes advantage of redundant information for photons which interact three or more times in the instrument (3+ site events) to determine the most probable interaction order. CKD additionally allows efficient rejection of background events, including photons which scatter out of the instrument before fully depositing their energy (Compton continuum photons), non-localized $\\beta^{-}$-decays, $\\beta^{+}$-decays, and pair-production events. The second reconstruction technique, Single Scatter Discrimination (SSD), allows good determination of the interaction order in 2-site events, but without the benefit of background rejection. (Also in Paper I we summarized additional techniques for rejecting $\\beta^{-}$-decays, $\\beta^{+}$-decays, and pair-production events.) In addition to CKD and SSD, other reconstruction techniques can be imagined. As we discussed in Paper I, for a given GCT configuration the performance will depend on the reconstruction techniques employed. Telescope performance will also depend on \\textit{event cuts}, which can be made on the initial direction of the photon scatter, the number of interaction sites, and the minimum separation between the first and second interaction sites (\\textit{minimum lever arm}). The tradeoffs are generally higher efficiency at the expense of degraded angular resolution, and hence increased background. First, the uncertainty in the Compton scatter angle (angular resolution) is smaller for forward scatter events than backscatter events (Paper I, Equation 4). Second, events with only 2 interaction sites (2-site) do not permit CKD background rejection, and also have a larger fraction of backscatter events that 3+ site events. Finally, a larger minimum lever arm will minimize the effects of spatial uncertainty in the detectors, improving angular resolution and hence background, but at the expense of lower efficiency. \\begin{figure} \\centering \\includegraphics[width=0.9\\linewidth]{figure2.eps} \\caption{Angular resolution for on-axis sources as a function of photon energy. Also shown is the limit set by Doppler broadening of bound electrons.\\label{figure3}} \\end{figure} Here we present detailed simulations of a GCT configuration in an effort to determine the optimized sensitivity to nuclear line emission, as well as demonstrate the variation in performance for several different combinations of reconstruction techniques and event cuts. Our selections range from utilizing most of the event information to physically reconstruct the event (Case 1, with CKD), to using a purely empirical, but highly efficient approach (Case 4), and should fairly represent the range of performance characteristics possible within a GCT. \\begin{figure} \\centering \\includegraphics[width=0.9\\linewidth]{figure3.eps} \\caption{Photopeak spectral resolutions, which show little dependence on the reconstruction technique or event cuts. For comparison is shown the assumed single-site resolution.\\label{figure4}} \\end{figure} ", "conclusions": "GCTs offer an attractive option for a soft $\\gamma$-ray observatory following INTEGRAL. The GCT presented here will allow a factor of 10 improvement in nuclear line sensitivity over INTEGRAL, which is required for new scientific goals such as the systematic study of Type Ia SNe, as well as improved imaging of the positron annihilation line, $^{26}$Al, $^{60}$Fe, and $^{44}$Ti. We have also demonstrated how a judicious choice of reconstruction techniques (CKD) and event cuts (3+ site, forward scatters) results in a 3-fold improvement in sensitivity near 1~MeV, while maintaining angular resolution near the Doppler-broadening limit. Our next goal is to study the performance of several different GCTs configurations, varying the detector-plane spacings, to determine how the geometry affects performance." }, "0010/astro-ph0010107_arXiv.txt": { "abstract": "The ability of the neutron star surface to supply all or only part of the charges filling the pulsar magnetosphere is crucial for the physics prevailing within it, with direct consequences for the possible formation of pair creation regions. We evaluate the conditions for $e^-$ emission from pulsar surfaces for a simple Goldreich-Julian geometry taking both thermal and field emission processes into account. Using recently published estimates for the equation of state at the neutron star's surface, we show, that for a large range of $T_{\\rm surf}, B$ and $P$, the liberated charges will fully screen the accelerating B-parallel electric field ${\\rm E_\\|}$. For surface temperatures $T_{\\rm surf}<2\\cdot 10^5 $K a balance between field emission of electrons and shielding of the field will occur. Even in the overidealised case of $T_{\\rm surf}=0$ one can expect a prodigious supply of electrons which will weaken the accelerating ${\\rm E_\\| }$. We calculated the motion of electrons along selected polar field lines numerically for the low temperature, field emission scenario yielding their Lorentz factors as well as the produced radiation densities. Inverse Compton and later curvature losses are seen to balance the acceleration by the residual electric fields. We found that the conditions for magnetic pair production are not met anywhere along the field lines up to a height of 1500 pulsar radii. We did not {\\it a priori} assume an \"inner gap\", and our calculations did not indicate the formation of one under realistic physical conditions without the introduction of further assumptions. ", "introduction": "A proper assessment of the conditions very close to the surface of a neutron star is a vital ingredient for any model that attempts to explain the enigmatic origin of the observed electromagnetic radiation from pulsars. Numerous studies of the polar cap acceleration have been made within a particular framework: A steady acceleration of electrons within a space charge limited particle flow originating at the surface of the neutron star will produce particles emitting magnetically convertible curvature $\\gamma-$rays (e.g. Arons \\& Scharlemann 1979). These high energy photons are converted into $e^+e^-$-pairs with high enough efficiency (Daugherty \\& Harding 1982) so that in the end, they and their daughter products will exceed the initial (Goldreich-Julian density) by a factor of $10^4$ (e.g. Zhang \\& Harding 2000). An important {\\it a priori } assumption is inherent in all these calculations: Above a critical height $h_{\\rm gap}$ a region of high conductivity will quench any electric field ${\\rm E_\\|}$ parallel to the guiding B-field lines, so that the full potential across the polar cap will be active between the outer surface of the neutron star and the inner surface of the high conductivity region. Our approach differs in that respect from other workers, we do not make that initial assumption, but try to calculate the amount of charges produced by the surface, how they affect and are affected by ${\\rm E_\\|}$, their energy losses and the possibility of pair production. Should one find suitable conditions for such a pair production region to exist, the methods quoted above will be the prime choice for its description. The conclusion that electron-positron pairs with high densities exist close to the neutron star has recently been questioned by several authors. Kunzl et al. (1998) considered this widespread assumption of dense pair plasmas for the problem of radio wave propagation within the neutron star magnetosphere. On hand of the most recent emission height estimates (50 - 100 $R_{\\rm NS}$) for the radio emission (Kijak \\& Gil (1997); Kramer et al. (1997, 1998)) they found that the existence of low frequency radio emission for example from the Crab pulsar is incompatible with the scenario of extended dense pair production. These findings were fully confirmed by the theoretical investigation of Melrose \\& Gedalin (1999), who treated the plasma emission mechanisms in detail and came to the conclusion that only low energy particles with Goldreich-Julian density could exist in the radio emission regions in order to alow for escaping radio waves (see also Melrose (2000) and references therein). Furthermore recent simulations of the pair cascade by Arendt \\& Eilek (2000) for strong and weak magnetic fields shed some light on the difficulties for an efficient production of secondary pairs by high energetic primary particles via $\\gamma$-ray curvature photons. Although their findings need some more investigation they could show that the field strength at the surface has the most important effect on the cascade. For strong fields $B\\sim 10^9\\,{\\rm T}$ they find almost no secondary photons and the cascade terminates in one generation. The weak fields ($B\\sim 10^7\\,{\\rm T}$) had either no cascades or quite weak ones. Only for moderate fields ($B\\sim 10^8\\,{\\rm T}$) they observed some secondary particle generations. Arendt \\& Eilek concluded that it is not yet clear whether pulsars can really produce a large number of pairs at the surface gaps. This tempted us to question the canonical {\\it a priori} assumption of the existence of an inner pair creating gap and investigate its possibility given a reasonable estimate of the surface characteristics and the fields involved. In the literature there are three fundamentally different ideas for pair production in the inner magnetosphere: A \"starved\" magnetosphere where the neutron star surface provides only an insufficient supply of charged particles (e.g. Ruderman and Sutherland (1975)), a magnetosphere where the Goldreich-Julian density is reached at the surface but due to curvature effects of the field lines deviations occur close to the surface (e.g. Arons \\& Scharlemann (1979), Barnard \\& Arons (1982)) and Shibata's model which predicts electric fields caused by {\\em overdense} regions, that means in the opposite sense than in the Arons model (Shibata, 1997). The latter two ideas suffer from serious weaknesses since it turned out that Arons type models do not work for slow pulsars (Beskin (2000) gives an upper limit of 0.1 to 0.3 s for the period). The Shibata model in turn neglects the back reaction of an overdense region which reduces the particle flow from inside. Therefore we will restrict our discussion to models of Ruderman-Sutherland type and fix the parameter range in which this kind of models is reasonable. For the scope of this article we consider a strong magnetic dipole anchored in a heavy ($1.4 M_{\\rm sun}$) sphere with a radius $r= 10\\,{\\rm km}$ that rotates twice per second ($\\Omega=2 \\pi \\cdot 2 s^{-1}$). Let the B-field have a strength of $|B_0|=10^{8}~{\\rm T}$ at the surface and for simplicity`s sake let the magnetic and the rotational axes be aligned\\footnote{We use spherical coordinates $(r,\\theta,\\phi)$ centered in the neutron star with $\\theta$ being the directional angle referred to the pole which coincides with the rotation axis.}. \\begin{equation} {\\rm B}(r,\\theta) = B_0\\cdot \\left( {r_{\\rm NS} \\over r}\\right) ^{-3} \\cdot (\\cos\\theta ,{1\\over 2}\\sin\\theta , 0) \\end{equation} Such objects have been discussed as idealised cases of radio pulsars by Goldreich \\& Julian (1969) and many authors since then. Of course one has to realise that the above mentioned idealisations limit the applicability of such simple models for the interpretation of observational facts from the growing number of pulsars that is being discovered. We will however try to extend the scope of these models by constraining the number of charged particles available and analysing the fate of individual charged particles (limiting ourselves to electrons and positrons) in the polar regions of such heavy rotating dipoles under the constraints given by some of the more recent observations. Although a pulsar magnetosphere is characterized by extreme physical conditions in terms of strong electric and magnetic fields the fundamental processes that can supply charges are reasonably well known from terrestial experiments and technical applications and can be extrapolated to the surface conditions of a neutron star. Such an extrapolation provides constraints on the initial charge density which is vital for any model of cascading particle creation and the radiation processes. ", "conclusions": "" }, "0010/astro-ph0010277_arXiv.txt": { "abstract": "The first VLBI images of H$_2$O maser emission in the Circinus Galaxy AGN show both an accretion disk and outflow 0.1 to 1 pc from the central engine. The maser traces a warped, edge-on accretion disk between radii of about 0.1 and 0.4 pc that is bound by a $1.3\\times10^6$ M$_\\odot$ central mass. The rotation curve is somewhat flatter than a Keplerian rotation law and is consistent with a disk mass on the order of $10^5$ M$_\\odot$. Clumpy substructure may display a kinematic signature consistent with spiral arms. Away from the disk, a second population of water masers traces broadly a bipolar, wide-angle outflow that contains (bullet-like) clumps ejected from a region centered on the central engine and $<0.1$ pc in radius. Out to a radius of $\\sim 0.4$ pc, the warp of the accretion disk appears to channel the outflow, the orientation of which on the sky coincides with the orientation of the known kiloparsec-scale ionization cone. Beyond this radius, the flow crosses the disk and truncates it. This suggests that the current accretion event and associated disk-outflow geometry have a lifetime on the order of $10^7$ years. ", "introduction": "The Circinus galaxy is one of the nearest Seyfert\\,II galaxies that hosts an extragalactic H$_2$O maser. (Freeman \\etal\\ 1977 estimate a distance of 4~Mpc.) The $\\sim 10^{42}$ erg s$^{-1}$ (2-10 keV) central engine is obscured at energies below 10 keV by a large gas column, n$_{\\rm H}\\sim4\\times10^{24}$ cm$^{-2}$ (\\cite{matt99}). The X-ray spectrum also exhibits a prominent Fe K$\\alpha$ line that probably comprises reflected radiation (\\cite{matt96}). These characteristics suggest some similarity to another maser-host galaxy, NGC\\,1068. The Circinus AGN exhibits an ionization cone (\\cite{marconi94}) within which Veilleux \\& Bland-Hawthorn (1997) observed linear optical filaments and compact knots reminiscent of Herbig-Haro objects. The cone opening angle is at least $90^\\circ$, at a mean position angle (PA) of roughly $290^\\circ$. The AGN also drives a kpc-scale nuclear outflow betrayed by bipolar radio lobes (\\cite{elmouttie98}) at PA $\\sim 295^\\circ$, which are largely aligned with the minor axis the galactic H\\,I disk (\\cite{jones99}) and a nuclear $^{12}$CO ring (\\cite{curran98}). There is strong evidence that at least several extragalactic H$_2$O masers trace edge-on accretion disks bound by central engines $\\ga 10^6$ M$_\\odot$. These masers display (1) roughly linear structure in projection on the sky, (2) emission symmetrically bracketting the galactic systemic velocities, and (3) declining rotation curves that indicate more or less Keplerian differential rotation. Discovery of similar symmetry in the Circinus H$_2$O maser spectrum (Nakai\\ \\etal\\ 1995; Greenhill\\ \\etal\\ 1997) led to speculation that there too the masers trace an accretion disk (\\cite{greenhill97}). Though the observation of an accretion disk in Circinus confirms that prediction, the discovery of qualitatively new, nondisk emission is equally interesting. In NGC\\,1068, disk emission is accompanied by secondary maser emission $\\gg1$ pc downstream in a jet, where material in the narrow line region deflects it (\\cite{gallimore96}). However, the nondisk emission in Circinus lies $<1$ pc from the central engine and is more closely related to it. ", "conclusions": "" }, "0010/astro-ph0010041_arXiv.txt": { "abstract": "A field of $\\sim 38^\\prime \\times 38^\\prime$ around the supernova remnant (SNR) G349.7+0.2 has been surveyed in the CO J=1--0 transition with the 12 Meter Telescope of the NRAO, using the On-The-Fly technique. The resolution of the observations is $54\\pp$. We have found that this remnant is interacting with a small CO cloud which, in turn, is part of a much larger molecular complex, which we call the ``Large CO Shell''. The Large CO Shell has a diameter of about 100 pc, an H$_2$ mass of 9.3$\\times 10^5$ M$_\\odot$ and a density of 35 \\c3 . We investigate the origin of this structure and suggest that an old supernova explosion ocurred about $4 \\times 10^6$ years ago, as a suitable hypothesis. Analyzing the interaction between G349.7+0.2 and the Large CO Shell, it is possible to determine that the shock front currently driven into the molecular gas is a non-dissociative shock (C-type), in agreement with the presence of OH 1720 MHz masers. The positional and kinematical coincidence among one of the CO clouds that constitute the Large CO Shell, an IRAS point-like source and an ultracompact H II region, indicate the presence of a recently formed star. We suggest that the formation of this star was triggered during the expansion of the Large CO Shell, and suggest the possibility that the same expansion also created the progenitor star of G349.7+0.2. The Large CO Shell would then be one of the few observational examples of supernova-induced star formation. ", "introduction": " ", "conclusions": "" }, "0010/astro-ph0010088_arXiv.txt": { "abstract": "We reveal the role of clustering of sub-clumps, which is expected in the cold dark matter (CDM) universe, in forming a bright elliptical galaxy (BEG) from a low-spin seed galaxy. This can be done by comparing the evolution of a low-spin seed galaxy including small-scale density fluctuations expected in the CDM universe (Model 1) with that of a completely uniform one (Model 2), using numerical experiments. We show that Model 2 cannot reproduce the properties of BEGs and forms a disk which is too compact and too bright due to the conservation of the initial-small angular momentum. In Model 1 clustering of the sub-clumps caused by initial small-scale density fluctuations leads to angular momentum transfer from the baryon component to the dark matter and consequently a nearly spherical system supported by random motions is formed. Moreover the collisionless property of the stars formed in the sub-clumps prevents the dissipative contraction of the system, leading to a large measured half-light radius. As a result, the end-product is quite well reproduces the observed properties of BEGs, such as the de Vaucouleurs light-profile, typical color and metallicity gradients, the large half-light radius, the small ratio of the rotational velocity to the velocity dispersion ($V/\\sigma$). We conclude that the clustering of sub-clumps, i.e., the hierarchical clustering, plays a crucial role in the formation of BEGs from a low-spin seed galaxy. ", "introduction": "Observed bright elliptical galaxies (BEGs) have the common interesting properties, which are expected to provide strong constraints on their formation scenario. For example, Davies et al.\\ (1983) found that the ratios of the rotation velocity to the velocity dispersion, $V/\\sigma$, of BEGs are much smaller than those of faint ellipticals. In photometric properties, the light profile of BEGs is well described by the de Vaucouleurs law, whereas faint ellipticals usually have an exponential light profile (e.g., Caon, Capaccioli, \\& D'Onofrio 1993). On the other hand, Kawata (1999) showed that in the CDM universe, i.e., the hierarchical clustering scenario, elliptical galaxies can be formed from a low-spin overdense region, using numerical simulation which includes almost all the important physical processes in galaxy formation self-consistently. He adopted a semi-cosmological model proposed by Katz \\& Gunn (1991). In this model, the process of galaxy formation is mimicked by a collapse of a top-hat over-dense sphere (a seed galaxy) which initially follows a Hubble flow expansion and has a solid-body rotation as an effect of the external tidal field. The amount of the initial solid body rotation is specified by a dimensionless spin parameter, $\\lambda$ (Peebles 1971). In addition, small-scale density fluctuations expected in the CDM universe are superposed on this sphere which consists of dark and baryonic matter. Using this model, Katz \\& Gunn (1991), Katz (1992), and Steinmetz \\& M\\\"uller (1994, 1995) studied only the case of $\\lambda=0.08$ and seemed to succeed in forming a system which has similar properties to an observed disk galaxy in several respects (see also Koda, Sofue, \\& Wada 2000a; 2000b, who presented, in the same way, that disk-like systems are formed for a range of $\\lambda=0.10-0.04$). However, the spin parameter of the virialized dark matter falls in a range 0.02--0.11 with a median value of 0.05, according to the results of N-body simulations (e.g., Barnes \\& Efstathiou 1987; Warren et al.\\ 1992) and analytical works (e.g., Steinmetz \\& Bartelmann 1995; Eisenstein \\& Loeb 1995), in which $\\lambda=0.08$ is near a high-end value. Kawata (1999) studied the evolution of a seed galaxy which has a slow rotation corresponding to $\\lambda=0.02$ and showed that the end-product reproduces the observed properties of elliptical galaxies very well. Cosmologically, a small spin-parameter is preferred in a galactic halo collapsing at a high redshift (Heavens \\& Peacock 1988). Moreover, elliptical galaxies are considered to be a system whose stellar component is old and rotates slowly. Thus, it is a quite natural consideration that most elliptical galaxies were formed in a halo collapsing at high redshift and rotating slowly. \\begin{figure*} \\epsscale{.9} \\plotone{f1.eps} \\figcaption[f1.eps] { Time evolution of the system in Model 1. The upper, middle, and lower panels show the distributions of the dark matter, gas, and stars respectively. Each panel measures 150 kpc across and shows the $x$-$z$ projection of the particles, where we set the $z$-axis to be the initial rotation axis. \\label{anim1}} \\end{figure*} Kawata's study was based on the hierarchical clustering scenario, because in his simulation the galaxy forms through mergers of sub-clumps caused by initial small-scale density fluctuations (see Fig.\\ 1 of Kawata 1999). The question we have to ask here is how the clustering of sub-clumps acts on the formation of elliptical galaxies from a low-spin seed galaxy. In order to understand further the role of the clustering, we compare the evolution of a low-spin seed galaxy including small-scale density perturbations with that of a uniform one, as the opposite case. Then, we focus on massive seed galaxies, and examine whether both types of seed galaxy can reproduce the observed properties of BEGs quantitatively, because the properties of BEGs, as mentioned above, are expected to be able to provide constraints on their formation scenario. If the uniform low-spin seed galaxy also succeeds in forming a system similar to BEGs, small-scale density fluctuations superposed on the initial slow-rotating sphere will not be indispensable to the model. The present study will thus clarify the role of the clustering of the sub-clumps induced by small-scale density fluctuations in galaxy formation from a massive low-spin seed galaxy. The plan for the remainder of the paper is as follows. In Section 2, the numerical method and the model are described. Section 3 presents results of numerical simulations. Conclusion is described in Section 4. \\begin{figure*} \\epsscale{.9} \\plotone{f2.eps} \\figcaption[f2.eps] { The same as Fig.\\ 1 but for Model 2. \\label{anim2}} \\end{figure*} ", "conclusions": "We showed that the luminosity profile, color and metallicity gradients, half-light radius, and $V/\\sigma$ of the end-product formed by clustering of sub-clumps are in very good agreement with those of the observed BEGs quantitatively. Since the CDM cosmology naturally predicts the existence of massive low-spin seed galaxies including small-scale density fluctuations, the CDM scenario is quite suitable for the formation of BEGs from a massive low-spin seed galaxy, in which the clustering of sub-clumps plays an essential role. On the other hand, in a massive low-spin seed galaxy including no initial small-scale density fluctuation, the small initial spin leads to a compact disk-like object dominated by the rotation owing to the angular momentum conservation. The disk has an exponential light profile, which is obviously inconsistent with the light profiles of the observed BEGs. Furthermore, the central surface brightness is too high to be in agreement with that of disk galaxies. No object like this has been observed in our universe despite its high surface brightness. This result suggest that a massive low-spin seed galaxy is not allowed to have a uniform density." }, "0010/astro-ph0010263_arXiv.txt": { "abstract": "Seventeen southern sky BL Lacs were observed in UBVRI using the CCD Camera on the 1.0m telescope at the South African Astronomical Observatory (SAAO) in Aug and Nov 1999. The analyses of all the seventeen sources are now complete, and are available via anonymous ftp (ftp pukrs1.puk.ac.za/pub/Blazars). A few examples of our results are however given in this paper. Whereas PKS 2005-489 and 2155-304 appear to have been in a high state, PKS 0048-097 and PKS 0521-365 showed evidence of variability on a time-scale of a few days, with the amplitude of variability increasing towards short wavelengths. This is consistent with observations of gamma-ray BL Lacs, which show similar behaviour in optical and X-rays. ", "introduction": "BL Lacs are the only members of the population of extragalactic sources which have been observed at TeV energies. They emit their energy over a broad range of the spectrum (radio to $\\gamma$-rays), with signatures of variability in optical, X-rays and $\\gamma$-rays, which are large compared to other AGNs. BL Lacs are highly polarized non-thermal sources with compact flat-spectrum radio cores. Most BL Lacs were identified in radio and X-ray surveys. Whereas those identified in X-rays (XBLs) are mostly nearby, the radio selected BL Lacs (RBLs) have a larger distance scale. Whereas a few XBLs were detected as TeV sources, we find that the RBLs were mostly detected as EGRET sources (see e.g. Lin et al. 1997, 1999). The number of BL Lacs known presently are mostly northern sources, which constitute $\\sim$75\\% of all known BL Lacs. Convincing detections of VHE photons from BL Lacs have mainly been made from Northern Hemisphere sources, which include Mrk 421 (Punch et al. 1992), Mrk 501 (Quinn et al. 1996) and 1ES 2344+514 (Catanese et al. 1998). TeV observations of Southern Sky sources mostly gave upper limits (e.g. Roberts et al. 1999). Besides PKS 2155-304, which has also been recently confirmed as a TeV source (Chadwick et al. 1999), none of the southern BL Lacs have been well monitored. Therefore, it is justifiable to claim that there are a number of BL Lacs that have not been identified from the southern skies, and probably with advances in VHE ground-based telescopes, some are likely to be identified as TeV sources in future. Our objective is to monitor Southern Sky BL Lacs in optical, and such information can be combined with contemporary X-ray observations for modeling purposes to identify further TeV candidates. ", "conclusions": "Whereas BL Lac variability is a common feature, our visual multicolor observations provide important information concerning the relative flux change vs. frequency. The observation of an increasing amplitude of flux variability towards higher frequencies (in this case for PKS 0048-097 and PKS 0521-365) should be combined with simultaneous X-ray observations. This synchrotron feature in the optical/X-ray region of the spectrum should also imply a similar behaviour in the inverse Compton ($\\gamma$-ray) component of the spectrum." }, "0010/astro-ph0010113_arXiv.txt": { "abstract": "The straylight contamination due to the Galactic emission (GSC, Galaxy Straylight Contamination) entering at large angles from the antenna centre direction may be one of the most critical sources of systematic effects in observations of the cosmic microwave background (CMB) anisotropies by future satellite missions like {\\sc Planck} and MAP. For the Low Frequency Instrument (LFI), this effect is expected to be particularly crucial at the lowest frequency channels. We describe here different methods to evaluate the impact of this effect and compare it with other systematics of instrumental and astrophysical origin. The results are presented in terms of simulated data streams and maps, Fourier series decomposition and angular power spectrum. The contributions within few degrees from the beam centre dominate the GSC near the Galaxy plane. The antenna sidelobes at intermediate and large angles from the beam centre dominate the GSC at medium and high Galactic latitudes. We find a GSC peak at $\\sim 15\\mu$K and a GSC angular power spectrum above that of the white noise for multipoles $\\ell \\lsim 10$ albeit smaller than that of CMB anisotropies by a factor larger than $\\sim 10$. At large multipoles, the GSC affects the determination of CMB angular power spectrum significantly less than other kinds of instrumental systematics, like main beam distortions and $1/f$ noise. Although the GSC is largest at low Galactic latitudes, the contamination produced by far pattern features at medium and high Galactic latitudes, peaking at $\\sim 4\\mu$K, has to be carefully investigated, because the combination of low amplitude of Galaxy emission in those regions with the extremely good nominal {\\sc Planck} sensitivity naturally makes high Galactic latitude areas the targets for unprecedentedly precise estimation of cosmological CMB anisotropy. This paper is based on {\\sc Planck} LFI activities. ", "introduction": "After the great success of \\COBE-DMR (Smoot \\etal 1992, Bennet \\etal 1996a, G\\'orski \\etal 1996) which probed the gravitational instability scenario for structure formation through the detection of CMB anisotropies at few degree scales, and the recent balloon-borne experiments at high sensitivity and resolution on limited sky regions (De~Bernardis \\etal 2000, Hanany \\etal 2000), supporting a universe model with $\\Omega_{tot} \\sim 1$ (Lange \\etal 2000, Balbi \\etal 2000, Jaffe \\etal 2000), ultimately, future progresses of the CMB anisotropy cosmology will be based on two space missions, MAP (Microwave Anisotropy Probe) (see Bennet \\etal 1996b) by NASA and {\\sc Planck} by ESA, planned to be launched respectively in the years 2001 and 2007. In particular, the Low Frequency Instrument (LFI, Mandolesi \\etal 1998) and the High Frequency Instrument (HFI, Puget \\etal 1998) on-board {\\sc Planck} will cover together a wide frequency range (30$\\div$900 GHz) which should significantly improve the accuracy of the subtraction of foreground contamination from the primordial CMB anisotropy, providing at the same time a gold mine of cosmological as well as astrophysical information (e.g. De~Zotti \\etal 1999a and references therein). To fully reach these scientific goals, great attention has to be devoted to properly reduce and/or subtract all the possible systematic effects. Detailed simulation codes have been developed and are continuously implemented to analyse the impact of several classes of instrumental effects related to the behaviour of the optics, instruments and environment for a wide set of possible scanning strategies (e.g. Burigana \\etal 1998, Delabrouille 1998, Maino \\etal 1999, Mandolesi \\etal 2000a). Ultimately this effort leads to the optimization of the mission design and the production of realistic data streams and simulated maps for data analysis tools as well as a realistic definition of {\\sc Planck} scientific performance. In particular, the behaviour of the {\\sc Planck} antenna patterns both at intermediate and large angles from the directions of beam centres have to be carefully considered. The requirement on the rejection of unwanted radiation coming from directions far from the optical axis (straylight) is stringent for {\\sc Planck} and does not pertain only the telescope itself, but the entire optical system, including also solar panels, shielding, thermal stability and focal assembly components. The primary sources of error for the LFI are those due to imperfect off-axis rejection by the optical system of radiation from the Sun, Earth, Moon, planets, Galaxy and the spacecraft itself (de~Maagt \\etal 1998). The variations of the spurious straylight signal during the mission introduce contaminations in the anisotropy measurements. The antenna response features at large angular scales from the beam centre (far sidelobes) are determined largely by diffraction and scattering from the edges of the mirrors and from nearby supporting structures. Therefore they can be reduced by decreasing the illumination at the edge of the primary, i.e. increasing the edge taper, defined as the ratio of the power per unit area incident on the centre of the mirror to that incident on the edge. Of course, the higher is edge taper, the lower is the sidelobe level and the straylight contamination. On the other hand, increasing the edge taper has a negative impact on the angular resolution for the fixed size of the primary mirror (e.g. Mandolesi \\etal 2000a). A trade off between angular resolution and straylight contamination has to be found. The main aim of this work is to evaluate the impact of the Galactic emission as a source of straylight for {\\sc Planck} LFI. We will then compare it with the effects generated by other kinds of systematics, the main beam distortion introduced by optical aberrations and the $1/f$ noise related to gain fluctuations in LFI radiometers, and with the astrophysical contamination from the Galaxy and the extragalactic sources in the main beam. At LFI frequencies, the Galaxy straylight contamination (GSC) is expected to be particularly crucial at the lowest frequencies, due to the increase of synchrotron absolute emission and anisotropies with the wavelength. For simplicity, we limit our analysis to the case of the 30~GHz channel, but the methods presented here can be extended to higher frequencies. In Sect.~\\ref{simul} we briefly describe the basic recipes of our simulation code, discussing the geometrical aspects relevant for the full sky convolution, the format of the optical simulation output performed by the ESA (de~Maagt et~al.~1998) that are adopted in the present work, the conversion from data streams to maps and the computation of the Fourier modes and of the angular power spectra, two different estimators of the GSC impact. We estimated the expected GSC on the basis of the antenna integrated response from angular regions at different angles from the beam centre, and on the level of Galaxy emission. In Sect.~\\ref{acc_test} we focus on the integration accuracy of our computations and test the consistency of the code by assuming simplified input maps and antenna patterns. The main results concerning the evaluation of the GSC are presented in Sect.~\\ref{results}. In Sect.~\\ref{comparison} they are compared with the effects introduced by other kinds of instrumental effects and several sources of astrophysical contamination. Finally, we draw our main conclusions in Sect.~\\ref{conclusions}. ", "conclusions": "\\label{conclusions} We have studied the impact of GSC on {\\sc Planck} observations at 30~GHz, by considering different and complementary evaluation approaches: absolute and relative quantification of the impact on scan circle data streams, Fourier decomposition of scan circle signals, computation of maps of GSC and evaluation of their angular power spectra. These different methods allow us to focus on different aspects of GSC. No relevant differences are found by varying the angle $\\alpha$ between the spin axis and the telescope line of sight from $80^{\\circ}$ to $90^{\\circ}$. Our simulations show that the GSC peaks at values of about $15 \\mu$K (a value comparable with the sensitivity per pixel), mainly due to the signal entering at few degrees from the beam centre. Such values are found in the regions quite close to the Galactic plane, where in any case the ``direct'' (i.e. observed by the main beam) contamination from the Galaxy prevents an accurate knowledge of CMB fluctuations, as it is next to impossible to remove the Galactic signals to accuracy better than $\\sim$~ \\%. These large contamination values, although critical for CMB anisotropy measurements near the Galactic plane, are not crucial for the determination of Galaxy emission, which is several order of magnitude larger. By considering all the pixels in the sky, the typical values of GSC are less than the 50\\% of the white noise sensitivity per pixel. The most crucial contamination derives from the signal entering in the far pattern, in spite of its peak values, of about 4$\\mu$K, smaller than those obtained for the intermediate pattern regions. In fact, although this effect does not seem to be very large in amplitude (indeed, being nominally subdominant to power spectrum of expected receiver noise) it does dominate the GSC at medium and high Galactic latitudes, which are critical regions for the extraction of the best quality results on CMB anisotropy. This could be also critical for the {\\sc Planck} polarization measurements which will take advantage from the two patches close to the ecliptic poles where the sensitivity will be several times better than the average, according to the scanning strategy and the feed array arrangement. As expected on the basis of the behaviour of Galaxy emission angular power spectrum, the GSC affects the determination of the CMB angular power spectrum mainly in the low multipole region and much less at large multipoles, particularly when compared with the other instrumental effects considered here, the main beam distortion and the $1/f$ noise. The additional noise introduced by the main beam distortion can be in principle subtracted in the data analysis, provided that the beam pattern is accurately reconstructed. Of course, a substantial improvement in the data analysis is necessary to jointly treat all the systematics, of instrumental and astrophysical origin. From the telescope design point of view, the best optimization of the edge taper requires a trade off between the main beam distortion and the GSC effects." }, "0010/astro-ph0010439_arXiv.txt": { "abstract": "We observed the energetic binary Cygnus X-3 in both quiescent and flaring states between 4 and 16 $\\mu$m using the {\\it ISO} satellite. We find that the quiescent source shows the thermal free-free spectrum typical of a hot, fast stellar wind, such as from a massive helium star. The quiescent mass-loss rate due to a spherically symmetric, non-accelerating wind is found to be in the range (0.4--2.9)$\\times 10^{-4}$ $\\rm M_{\\odot}$ yr$^{-1}$, consistent with other infrared and radio observations, but considerably larger than the $10^{-5}$ $\\rm M_{\\odot}$ yr$^{-1}$ deduced from both the orbital change and the X-ray column density. There is rapid, large amplitude flaring at 4.5 and 11.5 $\\mu$m at the same time as enhanced radio and X-ray activity, with the infrared spectrum apparently becoming flatter in the flaring state. We believe non-thermal processes are operating, perhaps along with enhanced thermal emission. ", "introduction": "The Infrared Space Observatory {\\it ISO} (Kessler et al.\\ 1996) with its on-board complement of cameras and spectrometers, provided a rare opportunity to study sources at wavelengths not accessible from the ground. We present photometry and images of the Cygnus X-3 field made using ISOCAM (Cesarsky et al.\\ 1996), the camera on {\\it ISO}, in the wavelength range 4 to 16 $\\mu$m. We report the first spectrum of this source in this wavelength range. Cygnus X-3 is an enigmatic and unusual X-ray binary which was discovered $>$30 years ago (Giacconi et al.\\ 1967). Observations of it have been made from radio to gamma rays and in every wavelength bands the system exhibits unusual behaviour. In X-rays and infrared, the emission is partially modulated by a 4.8 h period (Mason, C\\'{o}rdova \\& White 1986), and this is believed to be the orbital period of the system. In 1991, van Kerkwijk et al. observed Cyg X-3 in the infrared with the CGS4 spectrometer on UKIRT and discovered He\\,{\\sc i} and He\\,{\\sc ii} together with N emission lines which were modulated at the 4.8 h period (van Kerkwijk et al.\\ 1992; van Kerkwijk et al.\\ 1996). They concluded that the lines originated from the companion star, either a Wolf--Rayet of the WN7 class, or an earlier WN4--6 in an outburst phase. This was investigated in more detail by Fender, Hanson \\& Pooley (1999, hereafter FHP99), who took spectra during a variety of activity states and confirmed that the underlying quiescent spectral type is of an earlier WN4--5 Wolf--Rayet star. Galactic Wolf--Rayet stars have high velocity winds, up to 2500 km s$^{-1}$ (Crowther, Hillier \\& Smith 1995; Bohannan \\& Crowther 1999). The Wolf--Rayet wind from Cyg X-3 is believed to be the main contributor in providing mass and angular momentum loss which increase the orbital period ($\\dot{P} = -6.6 \\times 10^{-10}$ s s$^{-1}$, Kitamoto et al.\\ 1995). This period increase is consistent with a mass-loss rate of $10^{-6} M_{\\rm T}$ $\\rm M_{\\odot}$ yr $^{-1}$, where $M_{\\rm T}$ is the total mass in the system (Kitamoto et al.\\ 1995; Matz 1997). Kitamoto et al.\\ (1994) modelled changes in the radio activity, infrared and X-ray states caused by changes in the density of the WR wind (Table \\ref{wind}). The multi-wavelength predictions of Kitamoto et al.\\ (1994) were tested by FHP99, who find that during an outburst the infrared {\\it K} band spectrum becomes dominated by strong twin-peaked \\hbox{He\\,{\\sc i}} emission lines. These lines show strong {\\it V}/{\\it R} variability correlated with the orbital phase. Due to the small size of the binary ($\\sim 5\\;R_{\\odot}$ separation, 4.8 hour period) the observed line flux could not arise from a region of this size as the environment would be too hot. FHP99 conclude that the He\\,{\\sc i} emission lines are emitted from a region which is significantly larger than the binary separation, and that the variability of these lines indicate an asymmetric emission region, possibly from a flattened disc-like wind. \\begin{table} \\caption{Radio, infrared and X-ray variability based on a changing density wind, from Kitamoto et al.\\ (1994). The term `modulation' refers to the continuum. The states marked with an asterisk (*) have been expanded on and confirmed by Fender et al.\\ (FHP99).} \\begin{tabular}{llll}\\hline Regime \t\t& Emission \t& Rare wind \t& Dense wind \\\\ \\hline Radio\t\t& Continuum*\t& Quiescence\t& Flaring \\\\ Infrared\t& Continuum*\t& Weak\t\t& Strong \\\\ & Line strength*& Weak\t\t& Strong \\\\ & Ionisation*\t& High\t\t& Low\t\\\\ & Modulation\t& Large\t\t& Small\t\\\\ X-ray\t\t& Luminosity*\t& Low\t\t& High\t\\\\ & Spectrum\t& Hard\t\t& Soft \t\\\\ \\hline \\end{tabular} \\label{wind} \\end{table} The compact object in Cyg X-3 is less well defined. Cherepashchuk \\& Moffat (1994) suggest that the high X-ray luminosity (greater than the Eddington luminosity for a 1 ${\\rm M_{\\odot}}$ object) implies that the compact object is a black hole. Schmutz, Geballe \\& Schild (1996) related time-variations of infrared spectral lines to an orbital velocity and derived a mass for the compact object in the range of 7--40 $\\rm M_{\\sun}$, assuming realistic values for the mass of the Wolf--Rayet star and the inclination of the system. However, their model does not account for all the relative phasing of the X-rays and infrared lines. Ergma \\& Yungelson (1998) point out that if the progenitor to the binary is two massive helium stars then the mass-loss and common envelope phase of these two stars prevents the formation of a black hole \\& Wolf--Rayet binary with an orbital period less than several days. This ambiguity does not resolve itself if the compact object is a neutron star; Ergma \\& Yungelson (1998) also comment that the propeller motion of a spinning neutron star would prevent wind accretion in the system, thus removing the jet-creation mechanism. With a high mass, high mass-loss (wind) object in close proximity to a compact object, a high degree of interaction takes place. Radio observations at wavelengths ranging from 1--20 GHz have shown Cyg X-3 to undergo many states of emission and are used to define the state of Cyg X-3. For most of its time, the system occupies a {\\it quiescent} radio state, with 8.3 GHz flux around 100 mJy (Waltman et al.\\ 1994). {\\it Major flares} (8.3 GHz flux $>$ 1 Jy) occur at random intervals every 18 months or so and relaxation after a flare can last up to several weeks (Waltman et al.\\ 1995). More common are {\\it minor flares} (8.3 GHz flux 200--1000 mJy) and these usually occur during a flaring period typically several weeks duration, with several small flares occurring in close proximity to each other (Waltman et al.\\ 1996). Minor flares can reach peak emission in a few hours and decay over several hours (Ogley 1998). The minor and major flares have been shown to be intrinsically different in their production, rise and relaxation times (Newell 1996). Preceding a major flare, the system enters a quenching state with fluxes at 8.3 GHz dropping to around 10 mJy. The system returns to quiescence then flares (Waltman et al.\\ 1996). The main mechanism for radio flares is thought to be associated with the acceleration of relativistic electrons in the base of jets. Synchrotron emitting plasmons have been detected, travelling at $\\simeq$ 5 mas d$^{-1}$ on both arcsec and mas scales (Spencer et al.\\ 1986; Mioduszewski et al.\\ 1998). There is no confident determination of the velocity of ejecta: while a velocity of 0.3$c$ is generally taken as the speed on the larger arcsec scale (e.g. Spencer et al.\\ 1986), other highly-relativistic interpretations are possible on the mas scale (e.g. Ogley, Bell Burnell \\& Newell 1997; Newell, Garrett \\& Spencer 1998). The Wolf--Rayet contributes not only to the enigmatic qualities of the system, but also a significant and varying opacity. The true nature of the secondary is a top priority for the understanding of the system. Observations at the {\\it ISO} wavelengths provide a unique opportunity to study the Wolf--Rayet wind, and spectra obtained can track its true nature. ", "conclusions": "We report the first quiescent and flaring 4-16 $\\mu$m observations of Cygnus X-3. The quiescent spectrum is a continuation of the shorter wavelength IR spectrum, and fits well the WB75 model for thermal free-free emission from a wind. We obtain a mass loss rate, assuming spherical symmetry of $\\sim 10^{-4}$ $\\rm M_{\\odot}$ yr$^{-1}$, and note that while this is consistent with other determinations, it is an order of magnitude larger than both the mass loss rate required to derive the change in orbital period (assuming $M_{\\rm T}$ of the order $10\\;\\rm M_{\\odot}$), and the mass loss rate expected from the hydrogen column density to the X-ray source. These measurements are based on assumptions, which have been scrutinised carefully; while we can identify a number of areas where factors of a few can be gained, we have been unable to identify a single area where a factor of ten can be found. We agree with the conclusion of FHP99 that the wind is most likely spherically asymmetric, but if this is the only effect operating, large disc density enhancement factors are required. There are several phenomena which could cause $P_{\\rm orb}$ to be underestimated; we regard this determination of $\\dot{M}$ to be less reliable. During radio and X-ray outburst, the broadband infrared spectrum shows rapid variability and a spectral flattening associated with increased flux densities. This is in agreement with what is seen in the near-IR during rapid (minutes) flaring events (Fender et al.\\ 1996). Similar (but not identical) near-IR events observed from GRS 1915+105 appear to be associated with jet formation episodes. The emission mechanism responsible for this flat spectral component remains uncertain; both free-free and synchrotron emission have been proposed already in the literature, and await detailed studies to test their viability. Whatever the origin of this component, it seems likely that it is associated with hot gas in the jet from the system, which naturally explains its emergence during periods of enhanced radio (jet) activity." }, "0010/astro-ph0010169_arXiv.txt": { "abstract": "{A 0.242 $M_{\\sun}$ object that finally becomes a helium white dwarf is evolved from Roche lobe detachment down to very low luminosities. In doing so, we employ our stellar code to which we have added a set of routines that compute the effects due to gravitational settling, and chemical and thermal diffusion. Initial model is constructed by abstracting mass to a 1 $M_{\\sun}$ red giant branch model up to the moment at which the model begins to evolve bluewards. We find that element diffusion introduces noticeable changes in the internal structure of the star. In particular, models undergo {\\it three} thermonuclear flashes instead of one flash as we found with the standard treatment. This fact has a large impact on the total mass fraction of hydrogen left in the star at entering the final cooling track. As a result, at late stages of evolution models with diffusion are characterized by a much smaller nuclear energy release, and they evolve significantly faster compared to those found with the standard treatment. We find that models in which diffusion is considered predict evolutionary ages for the white dwarf companion to the millisecond pulsar PSR B1855+09 in good agreement with the spin-down age of the pulsar. \\endabstract ", "introduction": "\\label{sec_intro} Low mass, helium white dwarfs (WDs) have drawn the attention of many researchers who have devoted a large effort to their study. Currently, it is well established that helium WDs should have been formed in a binary system because an isolated star would reach such a configuration in a time-scale much longer than the present age of our Universe. Examples of recent studies on helium WD evolution are Althaus \\& Benvenuto (1997; 2000), Benvenuto \\& Althaus (1998), Hansen \\& Phinney (1998), Driebe et al. (1998) and Sarna, Antipova \\& Ergma (1999). A crucial issue in the study of helium WDs is the amount of hydrogen left atop the WD after the end of mass transfer episodes. Sarna et al. (1999) have presented detailed calculations of the pre - WD evolution during the mass transfer stage in a close binary system. They found that, after detachment, helium WDs are surrounded by massive hydrogen envelopes of 0.01 - 0.06 $ M_{\\sun}$ with a surface hydrogen abundance by mass of $X_{H}=$ 0.35 - 0.5. Massive hydrogen envelopes have also been found by Driebe et al. (1998), who have mimicked the binary evolution by abstracting mass from a $1 M_{\\sun}$ model at an adequate rate. Binary systems composed by a helium WD and a millisecond pulsar offer us a good opportunity to test the predictions of evolutionary calculations. Millisecond pulsars are thought to be recycled during the mass transfer stage (Bhattacharya \\& van den Heuvel 1991). When mass transfer ends, the pulsar begins to spin down. If we set zero ages at the end of such mass transfer, then the ages of the WD and the pulsar component should be the same. If the pulsar radiates as a dipole and its current period ($P$) is much longer than the initial one, then the spin-down of the pulsar is given by $t_{PSR}= P/2\\dot{P}$, where $\\dot{P}$ is the rate of change of $P$. Recently, van Kerkwijk et al. (2000) have optically detected the low-mass component of the PSR B1855+09 system and determined its effective temperature to be $4800 \\pm 800$ K. In addition, the mass of the WD is accurately known thanks to the measure of the Shapiro delay of the pulsar signal and results to be $0.258^{+0.028}_{-0.016} M_{\\sun}$. The age derived for the WD component results to be twice as long as the pulsar spin-down age ($\\approx 5 Gyr$) when using the models without diffusion of Driebe et al. (1998). Helium WD evolution in presence of element diffusion has been considered by Iben \\& Tutukov (1986), who computed the evolution of a $0.296 M_{\\sun}$ object throughout the entire evolutionary stages and found two thermonuclear flashes. Remarkably, they found that after flashes the amount of hydrogen in the envelope of the model was unable to support an appreciable amount of hydrogen burning reactions, enforcing a fast evolution for the resulting WD star. It is the purpose of this work to study the impact of element diffusion on the structure and evolution of a $0.242 M_{\\sun}$ helium WD star. This mass value is representative of the companion to PSR B1855+09. In doing so, two sequences are computed: one with and the other without diffusion. Thus, the resulting differences should be due to the allowance for such a process. ", "conclusions": "\\label{secc:conclu} From the results presented in the previous section, it is clear that element diffusion introduces important changes in the internal structure of the star. Our present detailed results indicate that, contrary to Driebe et al. (1998) and Sch\\\"onberner et al. (2000) assertions, nuclear reactions are not an important ingredient in determining the age of the star during the final cooling evolution. This conclusion is valid at least for the stellar mass value analyzed in this paper. Although the hydrogen mass fraction at the end of the computed stages is very similar in both sequences with and without diffusion, there is a large difference in the total time spent in reaching such a low luminosity in each case. While standard models without diffusion spend about 30 Gyr in reaching $\\log{L/L_{\\sun}}= -5$, models with diffusion get such a luminosity in about 15 Gyr. As stated in the introductory section, previous models available in the literature (those in which hydrogen burning is dominant during most of the cooling stages) lead to a strong discrepancy between the ages of PSR B1855+09 and its helium WD companion. These ages should be, on very general grounds, the same. The results presented in this paper strongly indicate that {\\it such a discrepancy is a direct consequence of ignoring element diffusion in the stellar model calculations}. If we include it properly, ages naturally come into a nice agreement. Consequently, we do not have to invoke any ad-hoc mass loss as proposed by Sch\\\"onberner et al. (2000) or exotic mechanisms to account for the observational situation of the binary system containing PSR B1855+09. It is worth noting that the structure of the models belonging to the final cooling branch we have found in the present paper for the 0.242 $M_{\\sun}$ helium WD strongly resemble those we have assumed in our previous work on helium WDs with hydrogen envelopes (Benvenuto \\& Althaus 1998). In fact, if we assume $M_{H}/M_{*}= 2 \\times 10^{-3}$ the agreement with the present results is excellent, not only regarding the cooling evolution but also with respect to the nuclear activity. In our opinion, the present detailed results largely justify the assumptions we have made in our quoted previous paper. We want to emphasize that our results strongly resemble those presented sometime ago by Iben \\& Tutukov (1986) (IT86) for the case of a $0.296 M_{\\sun}$ model. They found two flashes and suspected that the second one was due to diffusion, but because they did not compute models without diffusion, they were not conclusive about this point. Here we confirm that diffusion is the reason for the occurrence of additional flashes. IT86 also found that the final amount of hydrogen left in the WD is small enough to ultimately enforce a fast cooling. Although their model is more massive than ours, which prevents from a quantitative comparison between both set of calculations, the role of hydrogen burning at the final cooling stages is the same we have found. There is nevertheless an important difference between the work of IT86 and our study. In fact, IT86 considered that shortly after the second flash, when the model gets its maximum radius, it again overfills its Roche lobe and losses a tiny part of its stellar mass. Obviously, this mass belongs to the hydrogen envelope and such a loss has a large impact on the final value of $M_{H}/M_{*}$ in the WD. This is ultimately responsible for the fast cooling occurring at late stages of evolution. The point to emphasize here is that we did not consider such a mass transfer during our calculations, and some words about this point are in order. If such a transfer actually occurs, $M_{H}/M_{*}$ should drop below the values we presented above (eventually to zero). This would have an impact on the global structure of the WD. Such a situation may be considered in the frame of the models presented in Benvenuto \\& Althaus (1998). Apart from the structural differences we do not expect an important change in the time-scale of cooling. Notice that in the models presented in Benvenuto \\& Althaus, the time spent in reaching a given luminosity value is rather independent from the exact value of $M_{H}/M_{*}$ (if it is in the range of $1 \\times 10^{-3}$ to zero). Thus, the occurence of such a late mass transfer should not modify the conclusions about the age of the final WD model and its better agreement with the observations related to the ages of the PSR B1855+09 system." }, "0010/astro-ph0010443_arXiv.txt": { "abstract": "Quintessence models with a dark energy generated by pseudo Nambu--Goldstone bosons provide a natural framework in which to test the possibility that type Ia supernovae luminosity distance measurements are at least partially due to an evolution of the sources, since these models can have parameter values for which the expansion of the Universe is decelerating as well as values for which it is accelerating, while being spatially flat in all cases and allowing for a low density of clumped matter. The results of a recent investigation \\cite{NW} of current observational bounds which allow for SNe Ia source evolution are discussed. It is found that models with source evolution still favour cosmologies with an appreciable amount of acceleration in the recent past, but that the region of parameter space which is most favoured shifts significantly. ", "introduction": " ", "conclusions": "" }, "0010/astro-ph0010505_arXiv.txt": { "abstract": "We present the results of a wide-field survey for H$\\alpha$-emitting galaxies in the cluster AC\\,114 at $z=0.32$. Spectra centred on H$\\alpha$ at the cluster redshift have been obtained for 586 galaxies to $I_{tot}\\sim 22$ out to a radius of $\\sim 2\\,h_{50}^{-1}$\\,Mpc. At most, only $\\sim 10$\\% of these were found to be H$\\alpha$--emitting cluster members. These objects are predominantly blue and of late-type spiral morphology, consistent with them hosting star formation. However, $\\sim 65$\\% of the cluster members classified morphologically as spirals (with {\\it HST}), have no detectable H$\\alpha$ emission; star-formation and morphological evolution in cluster galaxies appear to be largely decoupled. Changes in the H$\\alpha$ detection rate and the strength of H$\\alpha$ emission with environment (as traced by local galaxy density) are found to be weak within the region studied. Star formation within the cluster members is also found to be strongly {\\it and} uniformly suppressed with the rates inferred from the H$\\alpha$ emission not exceeding 4\\,M$_{\\odot}$\\,yr$^{-1}$, and AC\\,114's H$\\alpha$ luminosity function being an order of magnitude below that observed for field galaxies at the same redshift. None of the galaxies detected have the high star formation rates associated with `starburst' galaxies; however, this may still be reconciliable with the known ($8\\pm 3$\\%) fraction of `post-starburst' galaxies within AC\\,114, given the poorly determined but short lifetimes of starbursts and the possibility that much of the associated star formation is obscured by dust. ", "introduction": "Clusters of galaxies provide a powerful laboratory for studying the evolution of galaxies. In the hierarchical picture of galaxy formation, clusters grow in mass by accreting galaxies from the surrounding field. Once these galaxies enter the dense environment of the cluster, star formation is suppressed and the galaxies evolve to become early-type galaxies. Many processes operate to make the cluster environment hostile to star forming galaxies. The most important processes include: mergers and violent encounters between galaxies (Lavery \\& Henry, 1994); tidal stripping and `harassment' (Byrd \\& Valtonen 1990, Moore et al.\\ 1996); ram pressure stripping of the gas disk (Gunn \\& Gott 1972, Abadi et al.\\ 1999, Quilis et al.\\ 2000); and simply the removal of any gas reservoir surrounding each galaxy (thus preventing replenishment of the gas disk; Larson, Tinsley \\& Caldwell 1980, Balogh et al.\\ 1999). Since clusters can be observed out to $z>1$, we can use the look-back times of the distant systems to directly investigate the evolution of cluster galaxies. Butcher \\& Oemler (1978a) and many subsequent authors have noted that the fraction of blue galaxies in cluster cores increases with redshift. Spectroscopic surveys have confirmed that these blue galaxies are indeed cluster members (e.g.\\ Dressler \\& Gunn 1982, 1983; Couch \\& Sharples 1987, hereafter CS; Lavery \\& Henry 1988; Dressler et al.\\ 1999) and have identified a conspicuous population of galaxies with no detectable emission (e.g.\\ [O{\\sc ii}]$\\lambda$3727) but with abnormally strong Balmer absorption lines (termed `E+A' galaxies by Dressler \\& Gunn, 1982 and `k+a' or `a+k' galaxies by Dressler et al.\\ 1999). It has also been proposed that the systematically higher blue galaxy fractions in distant clusters is coupled with a reduction in the S0 galaxy populations in these systems [identified from imaging with the {\\it Hubble Space Telescope} ({\\it HST})] in comparison to their local counterparts (Dressler et al.\\ 1997; Couch et al.\\ 1998, hereafter C98). The evolution of the galaxy populations in distant clusters may be explained as the result of the increased activity of field galaxies at intermediate redshift (Lilly et al.\\ 1996), combined with a higher accretion rate in these more distant systems (Bower 1991, Kauffmann 1996, Diaferio et al.\\ 2000): not only are the galaxies forming stars more rapidly when they arrive in the cluster, but they also arrive at a higher rate. However, while an overall picture of the evolution of cluster galaxies is emerging, we have no detailed understanding of the physical mechanism responsible for the suppression of star formation and its effect on other properties such as mass and morphology. The key issues can be summarised by three questions: (1)\\,By how much and how quickly is star formation suppressed in the infalling field galaxies? (2)\\,What is the radial dependence of the star formation suppression? (3)\\,Does the cluster simply suppress star formation or does it first promote a burst of star formation? The ability of current observations to address these issues is, however, seriously hindered by most distant cluster studies having largely concentrated on the central {\\it core} ($r<1\\,h_{50}^{-1}$\\,Mpc) regions in which the BO-effect was originally discovered: spectroscopic data of the quantity and quality needed to track in detail the star formation characteristics of galaxies well outside the core remain in short supply. Progress towards redressing this problem has been made by Balogh et al.\\ (1997, 1998), who used the CNOC Survey data (Yee, Ellingson \\& Carlberg 1996) to trace the spectral properties of galaxies out to and beyond the virial radius in 15 X-ray bright clusters at $0.18