{ "9803/astro-ph9803186_arXiv.txt": { "abstract": " ", "introduction": "Observations of elemental abundances in metal-poor halo stars provide important evidence regarding the early history, evolution and age of the Galaxy. Spectroscopic studies over a number of years have indicated the presence of neutron-capture, specifically rapid neutron-capture ({\\it i.e.} {\\it r}-process), elements in a number of these metal-poor halo stars (see {\\it e.g.} Spite and Spite 1978, Sneden and Parthasarathy 1983, Sneden and Pilachowski 1985, Gilroy et al. 1988, Gratton and Sneden 1991, 1994, Sneden et al. 1994, McWilliam et al. 1995a, 1995b, Cowan et al. 1996, Sneden et al. 1996, Burris et al. 1998). In addition, abundance comparisons of the {\\it r}-process and {\\it s}-process ({\\it i.e.} slow neutron-capture) elements between the oldest metal-poor halo stars and more metal-rich halo and disk stars provide direct evidence about the nature of the chemical evolution of the Galaxy. ", "conclusions": "Ground-based observations of the halo stars have indicated the presence of a number of neutron-capture elements. The availability of the HST has allowed for other spectral regions to be studied, and as a result more elements have been detected in these stars. For example, Figures 2 and 3 illustrate (see also Sneden et al. 1998) that we have now detected the element Ge in (three) halo stars. In addition, the 3$^{rd}$ {\\it r}-process peak elements, Os-Pt, as well as Pb have also now been detected in two stars using the HST (Cowan et al. 1996, Sneden et al. 1998). Including the ground-based detections of Th in CS~22892--052 (see Figure 1) and in HD 115444 (see Cowan et al. 1998), {\\it r}-process elements from proton numbers of Z = 32 to 90 have now been observed in metal-poor stars. This is a much wider range in proton (and mass) number than ever seen before and now includes the important 3$^{rd}$ {\\it r}-process peak. These detections further demonstrate that the {\\it r}-process, ranging up to the formation of the element Th, was in operation early in the history of the Galaxy. The observations also provide important information about the nature of the progenitors of the halo stars. The {\\it r}-process elements cannot be internally synthesized in the halo stars. Instead they must be produced in a previous generation (or generations) in an {\\it r}-process site. Since the metal-poor halo stars were formed early in the history of the Galaxy, presumably shortly after formation, the presence of {\\it r}-process elements in the halo stars requires very short evolutionary timescales for their progenitors. This further implies massive stars. While there is some uncertainty about the exact nature of the astrophysical site for the {\\it r}-process, it has long been suspected to be in supernovae, particularly core collapse supernovae from massive stars (see Cowan et al. 1991a). The abundance observations of the metal-poor halo stars appear to support that suspicion. It has also become clear with the accumulating data that the neutron-capture elements in the metal-poor stars have an abundance pattern that appears to be the same as the solar {\\it r}-process distribution. This is illustrated vividly in Figure 1, where all of the elements from Ba to Os in CS~22892--052 have relative solar abundances. While this correlation has been noted in the past (see {\\it e.g.} Gilroy et al. 1988), the high-resolution data in this star, covering a wide range of elements including Os in the 3$^{rd}$ {\\it r}-process peak, makes the argument much stronger. This same solar {\\it r}-process pattern also appears in other stars, as shown in Figure 2. Both ground-based and HST observations of HD~115444 ([Fe/H] = --2.7) show that the elemental abundance from Ba to Pt are consistent with a scaled solar {\\it r}-process curve. Further evidence of this is seen in the metal-poor halo star HD~122563 (Sneden et al. 1998). Elements such as Ba and La, which today are made predominantly in the {\\it s}-process, appear to have been made exclusively in the {\\it r}-process early in the history of the Galaxy. While this has been suggested previously (see {\\it e.g.} Truran 1981), the new observational data strongly support the contention that most (or all) of the elements were made in the {\\it r}-process at the earliest times in the Galaxy. The observations indicate the same relative {\\it r}-process abundance pattern in the oldest galactic stars and in the solar material, at least for elements with Z $\\ge$ 56. Therefore, the data indicate that the {\\it solar system {\\it r}-process abundances are not the result of global averages over different types of stars and epochs}. Instead, the stellar data suggest that the conditions that produced the {\\it r}-process elements are narrowly confined, perhaps both in terms of temperature and density of the nucleosynthesis and in terms of the mass range of the astrophysical sites (see Wheeler et al. 1998, Freiburghaus et al. 1998). The apparent lack of mixing early in the history of the Galaxy, when the relative abundance pattern is already apparent, demonstrated by Figure 6 also makes it less likely that the solar system {\\it r}-process distribution is the result of averages. We note in Figure 1, however, that while the elements in CS~22892--052 from Ba and above (Z $\\ge$ 56) are well-fit by the solar {\\it r}-process distribution, the extrapolation to the lower mass elements does not entirely fit the abundance data. In particular, while Sr and Zr do appear to be solar, the Y abundance is far below the solar curve. It is difficult to explain why two but not three of these neighboring elements in this star are solar. We note, further, that the abundance data for HD~115444, shown in Figure 2, show a similar separation between the lighter and heavier n-capture elemental abundances. In this star, again the abundances from Ba and above appear solar, but Sr-Zr do not. There may be several possible explanations. The weak {\\it s}-process, expected to occur during helium core burning in massive stars, is expected to contribute to the abundances of the elements from Sr-Zr. The data may be showing such a contribution and Cowan et al. (1995) even suggested some combination of the weak {\\it s}-process and the {\\it r}-process might be needed to explain the abundances of Sr-Zr in CS~22892--052. We note, however, one problem in this scenario is the apparent difficulty of producing {\\it s}-process elements in stars of extremely low metallicity. An alternative explanation may be that the more massive {\\it r}-process elements are synthesized in one site and the lower mass elements in another. Based upon meteorite data, Wasserburg et al. (1996) have suggested the existence of two {\\it r}-process sites with the separation in production occurring near mass number 140, {\\it i.e.} near Ba. Possible alternative {\\it r}-process sites have been discussed by Wheeler et al. (1998) and Baron et al. (1998). Further observations and analyses will be needed to understand the formation history of the lower mass {\\it r}-process elements. The most metal-rich of the halo stars studied here is HD~126238, with a metallicity of [Fe/H] = --1.7. We see the same basic trends for this star in Figure 3 that are seen in the more metal-poor stars CS~22892--052 and HD~115444. We note, however, that the abundance of Ba seems to lie above the solar {\\it r}-process curve. It was suggested by Cowan et al. (1996) that this might indicate some {\\it s}-process contribution to the Ba abundance. In other words, by the time that HD~126238 formed at a metallicity of --1.7, presumably more recently than the other two previously mentioned stars, some galactic {\\it s}-processing had occurred. It is seen in Figure 3 that the stellar Ba abundance is still below the total solar Ba abundance leading Cowan et al. to suggest that only the most massive stellar contributors to the {\\it s}-process had evolved at that point in time. Some support for their contention is given by Figure 4, which indicates the galactic chemical evolutionary trends for Ba as a function of metallicity. In this case metallicity is indicated by $\\alpha$, which may be a more reliable metallicity indicator than Fe, which is formed in both Type II and Type~I supernovae. These data spanning a wide range in metallicty might be explained by an evolutionary delay in the production of {\\it s}-process material. As demonstrated earlier in this paper, at early times in the Galaxy Ba apparently is produced from the {\\it r}-process. We note the clear change in slope in Figure 4 at a metallicity [$\\alpha$/H] $\\simeq$ --2 that appears to indicate the onset of the main {\\it s}-process nucleosynthesis production for Ba (and presumably other n-capture elements) in the Galaxy. Further evidence of this change in production mechanism, as a function of metallicity (and presumably time), for Ba is given in Figure 5. At very low metallicities (and early times) the {\\it s}-process element Ba and the {\\it r}-process element Eu appear to be synthesized solely in the {\\it r}-process. While there is scatter in the available data, we see that at the lowest values of [$\\alpha$/H] the [Ba/Eu] value in most of the stars is consistent with a pure {\\it r}-process origin. The long-lived radioactive nuclei (known as chronometers) in the uranium-thorium region are formed exclusively by the {\\it r}-process and can be used to determine the ages of stars and the Galaxy. One such chronometer, Th, has been detected in CS~22892--052 (Sneden et al. 1996, Cowan et al. 1997) (see Figure 1). Comparison between the initial abundance value produced in an {\\it r}-process site (often known as the production value) and the observed abundance value leads to a direct estimate of stellar ages. The abundances of the stable elements in the 3$^{rd}$ {\\it r}-process peak, a nuclear region nearby to the U-Th region, can be used to help constrain the predictions of the initial values of the long-lived radioactive chronometers, independent of knowing the site for the {\\it r}-process. Comparing the solar and the observed Th/Eu ratio in CS 22892--052, Cowan et al. (1997) found an age estimate of 15 $\\pm$ 4 Gyr for this star. They noted that consideration of galactic chemical evolution could lead to an older age of 17 $\\pm$ 4 Gyr. Pfeiffer et al. (1997) employed newer and more accurate nuclear data in the context of a waiting point approximation {\\it r}-process model. Using the stable stellar and solar data to constrain the predicted abundances of the radioactive {\\it r}-process nuclei, they compared the initial (as opposed to solar) value of Th/Eu with the stellar value and found a best estimate for this star of 13.5 Gyr. Their result for CS~22892--052 was not only consistent with Cowan et al. (1997), but is also consistent with recent globular cluster age determinations based upon Hipparcos data (see Pont et al. 1998). (See also Cowan et al. 1991a,b for a discussion of galactic and cosmological age determinations.) This technique, based upon predicted and observed radioactive chronometers, has been extended to an additional star with an age result approximately the same as that for CS~22892--052 (Cowan et al. 1998). We caution, however, that there are still many uncertainties, and to improve the accuracy of the chronometric estimates will require more observational and theoretical studies. It is encouraging to note, though, that the detection of thorium in CS 22892--052, and other halo stars, offers promise as an independent technique for determining stellar ages, and thus putting limits on galactic and cosmological age estimates." }, "9803/gr-qc9803087_arXiv.txt": { "abstract": "Despite the fact that the Schwarzschild and Kerr solutions for the Einstein equations, when written in standard Schwarzschild and Boyer-Lindquist coordinates, present coordinate singularities, all numerical studies of accretion flows onto collapsed objects have been widely using them over the years. This approach introduces conceptual and practical complications in places where a smooth solution should be guaranteed, i.e., at the gravitational radius. In the present paper, we propose an alternative way of solving the general relativistic hydrodynamic equations in background (fixed) black hole spacetimes. We identify classes of coordinates in which the (possibly rotating) black hole metric is free of coordinate singularities at the horizon, independent of time, and admits a spacelike decomposition. In the spherically symmetric, non-rotating case, we re-derive exact solutions for dust and perfect fluid accretion in Eddington-Finkelstein coordinates, and compare with numerical hydrodynamic integrations. We perform representative axisymmetric computations. These demonstrations suggest that the use of those coordinate systems carries significant improvements over the standard approach, especially for higher dimensional studies. ", "introduction": "\\label{sec:introduction} It is well known that the black hole solutions of the field equations of general relativity, with the Schwarzschild and Kerr solutions being astrophysically the most relevant, exhibit coordinate singularities when written in coordinates adapted to the exterior region. The very notion of a black hole was greatly clarified with the discovery, in the early sixties, of coordinate systems that remove those singularities, and indeed cover the whole spacetime~\\cite{Kruskal60} (for a general overview see~\\cite{MTW73},\\cite{Hawking}). The (rotating) black hole solution for the metric tensor, expressed in standard Boyer-Lindquist coordinates $(t,r,\\theta,\\phi)$ is singular at the roots of the equation $\\Delta=r^2-2Mr+a^2=0$ (where $M$ is the mass and $a$ the angular momentum aspect of the hole~\\cite{units}). In the spherically symmetric Schwarzschild case, the singularity at $r=2M$ is removable with the use of appropriate transformations of the radial and time coordinates. Slightly more complicated transformations, involving also the azimuthal coordinate, can remove the singularities at $r=r_{\\pm}=M \\pm (M^2-a^2)^{1/2}$ of a rotating black hole. The location of the coordinate singularity coincides with the event horizon of the black hole. Asymptotic observers lose causal contact with events near the black hole at precisely this location. Hence, despite the singular appearance of the metric, in simulations of matter flows in the background gravitational field of a black hole, one is in principle allowed to consider the open interval extending from the event horizon to some ``far zone\" at a large, but finite, distance from the hole. This approach has been widely used over the years in the numerical simulations of flows around black holes~\\cite{Wi72} -\\cite{Font98a}. The blow-up of the metric components at the horizon has implications on the behavior of hydrodynamical quantities. The coordinate flow velocity becomes ultra-relativistic and reaches the speed of light at the horizon. As a consequence, the Lorentz factor becomes infinite causing any numerical code to crash. Placing the inner boundary close to the horizon, required for capturing the effects of the relativistic potential, introduces large gradients in all hydrodynamical variables. The steep radial behavior makes the task of numerical evolution with a reasonable degree of accuracy challenging. Besides practical considerations, the approach of evolving only the exterior domain lacks of a well defined location for the inner boundary: the computation cannot include the horizon surface, but must commence at a location ``sufficiently close'' to it. Physically, the influence of the horizon region on the solution will progressively red-shift away the closer one gets to the horizon. The validity of a certain choice though, must be continuously reasserted, as new flows or effects are being investigated, with careful tests of convergence, as the inner boundary is progressively moved inwards. Such tests are complicated by the fact that, as mentioned in the previous paragraph, the solutions appear singular at the horizon and hence demand increasingly more resolution. Ameliorating those problems has motivated the use of a logarithmic radial coordinate (the so-called {\\it tortoise} coordinate~\\cite{RW57}). This technique relegates the event horizon to a infinite, negative, coordinate distance. An equidistant grid in the tortoise $r_{*}$ coordinate maps into an increasingly dense grid in the Schwarzschild $r$ coordinate, with infinite density at the horizon. This approach has proven successful in the extensive semi-analytic studies of black hole perturbation theory, and recently also for the axisymmetric integration of curvature perturbations as an initial value problem~\\cite{Krivanetal}. In wave systems, the ambiguity of the location of the inner boundary is addressed by the simple limiting form of the governing equations near the horizon. This is the well known fact that black holes act as finite potential barriers to electromagnetic and gravitational perturbations~\\cite{chandra}. The use of a tortoise coordinate does not bring similarly extensive benefits to the study of the hydrodynamical equations. The issue of the inner boundary location is less transparent for those equations. The steepness of the solution and the ``artificially'' high coordinate velocities persist, although they are now more treatable due to the substantial increase in resolution. In three dimensional simulations using Cartesian coordinates, the tortoise technique is not possible at all. It has been shown recently~\\cite{Papadopoulos98a} that {\\em adaptive mesh refinement} can indeed provide, even for 3D systems, the required resolution close to the event horizon. However, as alluded to above, these undesired pathologies can be eliminated in the case of fixed given black hole backgrounds with rather simple coordinate transformations. This reserves the power of adaptive mesh refinement for the more physically interesting features of the solution. There is considerable freedom in choosing coordinates regular at the horizon, which can be productively reduced by imposing criteria that can enhance their suitability for numerical applications. The obvious first criterion is of course the {\\em regularity of the metric}, in particular at the horizon. A second condition is that the constant time surfaces are {\\em everywhere spacelike}, as this is, currently, a pre-requisite for the implementation of modern numerical methods for relativistic hydrodynamic flows. An important third criterion is the {\\em time-independence} of the metric components. This leads to constant in time coefficients in the equations and simplifies disentangling the true hydrodynamical evolution from coordinate effects in the black hole background. Interestingly, we will see that those conditions still do not fix the coordinate system uniquely. We show that the number of available choices greatly reduces, but is still infinite, in the rotating black hole case. We give several examples of such systems, which we collectively call {\\em horizon adapted coordinate systems}. Such coordinate systems address the issues raised above in a straight-forward way: any radius {\\em inside} the horizon is equally appropriate (in the idealized continuum limit) for the imposition of a boundary condition, as the domain is causally disconnected from the exterior. Importantly, the irrelevance of the inner boundary location will persist even after the inclusion of other possible local physical processes that may be considered in conjuction with the hydrodynamical flow, e.g., radiative processes. The coordinate velocities of the flow will be bounded at the horizon, as they represent projections of the (finite) fluid four velocity onto a regular coordinate system. Hence the hydrodynamical nature of the flow becomes considerably less demanding on the integration algorithm. Gradients in the solution for {\\em scalar variables} such as pressure and enthalpy will of course persist. Those are physical and due to the curvature of the black hole, which requires a significant dynamic range for its resolution. The organization of this paper is as follows. In section II we introduce a class of horizon adapted coordinate systems for a non-rotating black hole. The rotating black hole and the more restricted class of coordinates available in that case are discussed in the Appendix. We outline the numerical hydrodynamical framework of our computations in section III. Exact solutions for spherical (Bondi) accretion are presented in section IV for both dust and perfect fluids. Section V describes the numerical results. In our numerical simulations we focus mainly on the spherically symmetric case, which captures the essential nature of the problem. Some axisymmetric computations are also briefly considered. The coordinate system on which we base our computations is the celebrated Eddington-Finkelstein coordinate system~\\cite{Eddington24},\\cite{Finkel58}. Our main aim in this report is to show the functionality of this class of coordinate systems as frameworks for the integration of the equations of relativistic hydrodynamics in black hole spacetimes. In three-dimensions (and rotating holes) computations of accretion onto black holes are likely to benefit significantly by the adoption of a horizon adapted coordinate system. ", "conclusions": "We have presented a family of {\\it horizon adapted coordinate systems} for the numerical study of accretion flows around black holes. In the rotating case we identify a discrete but infinite family, of which the first simple members are well known stationary coordinate systems. In the non-rotating case the freedom in building the regular stationary foliation is quantified by (at least) the space of bounded positive functions of one variable. We have shown how these systems allow for a better numerical treatment of accretion scenarios. Existing numerical studies of accretion flows onto black holes have been performed in the original, singular systems, i.e., Schwarzschild coordinates for the Schwarzschild solution and Boyer-Lindquist coordinates for the rotating (Kerr) solution. Although it is possible to solve the problem in these pathological coordinates, one is introducing artificial complexity, being forced to use very high resolution to deal with the unphysically large gradients that develop in the vicinity of the horizon. This may prevent the accurate solution of three dimensional problems. At the same time, the ambiguity regarding the position of the inner boundary of the domain (which should be the horizon) introduces a convergence criterion that must be enforced at all times if the solutions are to be trusted. We have focussed on the particular case of the Eddington-Finkelstein form of the Schwarzschild metric. The general relativistic hydrodynamic equations are now regular at the horizon, which permits an accurate description of accretion flows. We have demonstrated the feasibility of this approach with the numerical study of the spherical accretion (Bondi accretion) of dust and perfect fluid, and the comparison with the exact solutions which we re-derived in this coordinate system. We have also shown the functionality of the new coordinates in axisymmetric computations of relativistic Bondi-Hoyle accretion flows. In a forthcoming paper we plan to extend this approach to the rotating case, considering horizon adapted coordinate systems to study accretion flows in stationary Kerr spacetimes. Three dimensional accretion flows onto black holes are interesting both from an astrophysical and geometrical point of view, as they are thought to correspond to observable electromagnetic emission, and hence may help map the relativistic rotating black hole potential. The framework proposed here will help detailed numerical studies of such systems in the near future." }, "9803/astro-ph9803073_arXiv.txt": { "abstract": "The depletion of lithium during the pre-main sequence and main sequence phases of stellar evolution plays a crucial role in the comparison of the predictions of big bang nucleosynthesis with the abundances observed in halo stars. Previous work has indicated a wide range of possible depletion factors, ranging from minimal in standard (non-rotating) stellar models to as much as an order of magnitude in models which include rotational mixing. Recent progress in the study of the angular momentum evolution of low mass stars (Krishnamurthi \\etal 1997a) permits the construction of theoretical models which reproduce the angular momentum evolution of low mass open cluster stars. The distribution of initial angular momenta can be inferred from stellar rotation data in young open clusters. In this paper we report on the application of these models to the study of lithium depletion in main sequence halo stars. A range of initial angular momenta produces a range of lithium depletion factors on the main sequence. Using the distribution of initial conditions inferred from young open clusters leads to a well-defined halo lithium plateau with modest scatter and a small population of outliers. The mass dependent angular momentum loss law inferred from open cluster studies produces a nearly flat plateau, unlike previous models which exhibited a downwards curvature for hotter temperatures in the \\7li - T$_{\\rm eff}$ plane. The overall depletion factor for the plateau stars is sensitive primarily to the solar initial angular momentum used in the calibration for the mixing diffusion coefficients. The \\6li/\\7li depletion ratio is also examined. We find that the dispersion in the plateau and the \\6li/\\7li depletion ratio scale with the absolute \\7li depletion in the plateau and we use observational data to set bounds on the \\7li depletion in main sequence halo stars. A maximum of 0.4 dex depletion is set by the observed dispersion and \\6li/\\7li depletion ratio and a minimum of 0.2 dex depletion is required by both the presence of highly overdepleted halo stars and consistency with the solar and open cluster \\7li data. The cosmological implications of these bounds on the primordial abundance of \\7li are discussed. ", "introduction": "The status of big bang nucleosynthesis (BBN) as a cornerstone of the hot big bang cosmology rests on the agreement between the theoretical predictions and the primordial abundances of the light elements deuterium (D), helium-3 (\\3he), helium-4 (\\4he), and lithium-7 (\\7li) inferred from observational data (\\cite{YTSSO}; \\cite{WSSOK}; \\cite{CSTI}). Comparisons of this type often rely on models of galactic chemical and/or stellar evolution in order to associate the observed abundances of these light elements at the present epoch with their predicted primordial values. Recently the confrontation between prediction and observation has come under stricter scrutiny as it appeared that the primordial abundance of deuterium inferred from ISM/Solar data was smaller than its predicted abundance - the value of which follows from requiring that the standard big bang model predictions for \\4he are in good agreement with the abundance inferred from observations of metal-poor extragalactic \\hii regions (\\cite{Hata95}; \\cite{CSTII}; \\cite{Hata97}). Very recent measurements of the deuterium abundance along the lines-of-sight to high-redshift QSOs(\\cite{BT97}), along with a reanalysis of the \\4he data in light of new observations(\\cite{OSS96}), increase the tension between prediction and observation. The role of \\7li in these comparisons has been minor due to the large uncertainty in the estimate of the amount of lithium destruction during the lifetimes of \\popii (\\ie, metal poor) halo stars which could accomodate primordial lithium abundances ranging from the observed plateau value\\footnote{The lithium abundances of \\popii stars with $T_{\\rm eff} > 5800$K and [Fe/H] $< -1.3$ is nearly independent of metallicity ($[Li] = 2.25\\pm 0.1$, where $[X] = 12 + \\log y_X$ with $y_X$ the number ratio of X to hydrogen) and, hence, is referred to as a ``plateau\".} (no depletion) up to a factor of ten larger. This range in possible depletion factors is equally compatible with primordial lithium abundances which correspond to either ``low deuterium'' (which favors a primordial lithium abundance a factor of 3 higher than the plateau value), or the observed \\4he (which favors the plateau value). The arguments in favor of minimal lithium depletion are the flatness of the \\popii lithium abundance plateau (at low metallicity and high temperature) with respect to both metallicity and temperature and the low dispersion in the lithium abundance at fixed T$_{\\rm eff}$. This was generally consistent with ``standard'' stellar models (\\ie, models without rotation which burn lithium via convection during pre-main sequence evolution) which predicted little depletion (\\cite{DDKKR}). There are however some specific areas of disagreement in the comparison of the halo star data with standard models. The observed dispersion may be greater than that predicted (\\cite{DPD93}; \\cite{Thorburn94}; \\cite{Ryan96}) or not (\\cite{MPB95}; \\cite{Spite96}). There is a small population of highly overdepleted stars which appear normal except for \\7li (\\cite{Thorburn94}; \\cite{NRBD}), and the trends with metal abundance appear to conflict with expectations from the models (Thorburn 1994, \\cite{CD94}). However, the overall agreement between standard stellar evolution theory and observations in halo stars is good enough that nonstandard models would probably not be invoked to explain this data set. Therefore many investigators have adopted the reasonable assumption that the observed \\popii lithium abundances are close to the primordial value. The overall properties of \\7li depletion in standard models have been extensively studied (for a review see \\cite{MP97}) and there are some model independent predictions: There is partial \\7li depletion in the pre-main sequence (pre-MS), little or no main sequence (MS) depletion for stars hotter than about 5500 K, and there should be little or no dispersion in the \\7li abundance at a given T$_{\\rm eff}$ in clusters. Unfortunately, the observational data obtained from halo stars does not serve as a good diagnostic for these models since the initial abundance is not known and we have no information on the history of the \\7li abundance. Instead, one can look to the Sun and open clusters, systems with a nearly uniform initial lithium abundance ([Li] = 3.2 to 3.4) where detailed abundance data is available as a function of mass, age, and composition. These \\popi data provide stringent tests of theoretical models and it has been known for quite some time (\\eg, see \\cite{WS65} and \\cite{Z72}) that standard models fail these tests. The disagreement between the data and standard models has increased as both the observational data and the standard theoretical models have improved. The open cluster data provide clear evidence that the \\7li abundance decreases with increasing age on the MS, contrary to the standard model prediction that the convective zones of these hot ($\\geq$ 5500K) stars are not deep enough to destroy \\7li on the MS. In addition, there is strong evidence for a dispersion in abundance at fixed T$_{\\rm eff}$ and unexpected mass-dependent effects, such as the strong depletion seen in mid-F stars relative to both hotter and cooler stars (the ``Li dip'' of \\cite{BT86} - for extensive reviews of this issue, see \\cite{PH88}, \\cite{MC91}, \\cite{S95}, \\cite{Bal95}, and \\cite{MP97}). For standard models, these features cannot be explained by variations in the input physics (\\eg, opacities); rather, they indicate the operation of physical processes not usually included in standard models. By extension, these processes could also be operating in the \\popii stars. Another potentially useful diagnostic of lithium depletion can be found in globular clusters. Clusters provide samples which are homogeneous in age and composition, so one would expect a smaller dispersion in a cluster sample than in a field star sample. This is certainly the case for \\popi stars: \\popi field stars do show a larger dispersion than open cluster stars (\\cite{LHE91}; \\cite{FMS96}). However, recent observations of Li in globular cluster subgiant and turnoff stars show the opposite trend and create some serious difficulties for the minimal \\7li depletion scenario for metal poor stars (\\cite{DBK95}, \\cite{BDSK97}; \\cite{TDRRO97}). There are clear star-to-star differences in M92; out of 3 stars observed with very similar color, 2 have lithium abundances well below the plateau value. In NGC 6397, 20 stars near the turnoff were observed. For 7 with identical B-V color, there is a scatter of a factor of 2-3 in lithium abundance. Therefore a standard model treatment of the halo field stars must explain why the lithium abundances in globular clusters, but not in field stars, are anomalous. The disagreement between the \\popi lithium abundances and the predictions of standard stellar models has stimulated investigation of nonstandard stellar models. The most prominent explanations are mixing (either from rotation or waves), microscopic diffusion, and mass loss (\\cite{MP97}; \\cite{MC91}). For \\popi G and K stars, microscopic diffusion and mass loss are not likely explanations (\\cite{MC91}, \\cite{SF92}), which leaves mixing as a logical candidate. Mixing induced by rotation can explain many of the overall properties of the \\popi data. Rotational mixing can, on relatively long timescales, mix material from the base of the convective zone to interior regions of the star where lithium can be burned; both the rotation rate and the \\7li depletion rate decrease with increased age; and a distribution of initial rotation rates can produce a distribution of \\7li depletions and thus a dispersion in abundance at fixed T$_{\\rm eff}$. Although models with rotational mixing have the qualitative properties needed to explain the Pop I and Pop II data, two classes of objections have been raised: 1) the uniqueness of the solutions and adequacy of the physical model; and 2) discrepancies in the quantitative comparison of observation and theory. Rotational models require an understanding of the angular momentum as a function of radius and as a function of time. There have been persistent difficulties in reproducing the surface rotation rates as a function of mass and time in open clusters (\\cite{CDP95a},b; \\cite{KMC95}; \\cite{Bou95}). In addition, models with internal angular momentum transport from hydrodynamic mechanisms, such as the ones used in this paper, predict more rapid core rotation in the Sun than is compatible with helioseismology data (\\cite{CDP95a}, Krishnamurthi et al. 1997a (KPBS), \\cite{TST95}). Either of these difficulties could potentially affect the degree of mixing in the models. Models with rotational mixing predict that a range of initial angular momenta will generate a range of lithium depletion rates and therefore a dispersion in abundance among stars of the same mass, composition, and age. However, the difficulty in reproducing the angular momentum evolution of low mass stars has made quantifying the predicted dispersion difficult, and therefore only qualitative estimates have been made (\\eg, see \\cite{PKD}(PKD), \\cite{CD94}, \\cite{CDP95b}, and \\cite{CVZ92}). There have also been mismatches between the mean trend inferred from the data and from all classes of theoretical models. Chaboyer \\& Demarque (1994), for example, concluded that the observed mean trend of lithium abundance with T$_{\\rm eff}$ in metal-poor halo stars was not reproduced in {\\it any} class of models, be they standard, rotational, or with diffusion. There are also cases, such as the cool stars in young open clusters, where the observed dispersion is large but the rotational models predict little, if any, dispersion. The case for or against standard model lithium depletion is not as clear cut when we consider halo stars. It is argued that models with significant rotationally induced depletion could not produce a flat plateau with limited dispersion about the mean plateau abundance (\\eg, \\cite{BM97}). Furthermore, it was expected that such models would destroy far too much \\6li (\\cite{SFOSW}) in conflict with the observation of \\6li in HD 84937 (see section 4.2). On the other hand, standard (convective burning) models show lithium depletion trends contrary to the \\popi data. At minimum the mechanisms responsible for the \\popi \\7li pattern need to be identified and shown to not affect \\popii stars. It is difficult to construct a model where the nonstandard effects completely cancel for \\popii stars while still retaining consistency with the \\popi data. Models with microscopic diffusion predict modest \\7li depletion in plateau halo stars. In general these same models predict that the timescale for changes in the surface \\7li abundance decreases with increased mass; this produces a downward curvature in the \\7li-T$_{\\rm eff}$ relationship at high temperatures which is not observed. Mass loss can counteract this trend (\\cite{Sw95}; \\cite{VC95}). In either case, the net effect is modest depletion at the 0.2 dex level. Rotational mixing can also suppress diffusion (Chaboyer \\& Demarque 1994, \\cite{VC95}). The cancellation of different effects still requires at least some mean \\7li depletion in the halo stars. In a recent preprint \\cite{VC98} note that although the surface \\7li abundance varies strongly with mass in diffusive models, the peak subsurface abundance does not (note that the height of the peak is not preserved in models which include mixing). They then use the height of that peak to constrain the absolute \\7li abundance, arguing that the appropriate mass loss strips each star down to the region that contains this peak \\7li abundance. The survival of \\6li is also problematic in unmixed models with sufficient mass loss to expose the peak abundance; in a model of ours which can be compared to Figure 1 in \\cite{VC98}, \\6li drops to half its surface value in the outer 0.01 solar masses, well below the comparable mass content in the \\7li preserving region of 0.02 solar masses and also the peak in the \\cite{VC98} model. Further, it is disturbing to use a class of models which do not include a fundamental characteristic of stars, namely, rotation, when that property has been shown to be capable of affecting the issue being studied. In our view the most serious source of uncertainty in models with rotational mixing has been the understanding of the angular momentum evolution of low mass stars. We will show that many of the difficulties in reconciling observation and theory are resolved when models that are consistent with the rotation data in open clusters are used. In order to make definitive predictions of the amount of rotational mixing, it is necessary to follow the stellar angular momentum histories which requires knowledge of the initial distribution of rotation velocities along with an angular momentum loss law. There has been significant recent progress in constructing theoretical models consistent with the rotation data in young open clusters (KPBS; \\cite{CL94}; \\cite{KMC95}; \\cite{BFA97}; \\cite{A97}). With the latest generation of models, we can both infer the distribution of initial conditions and place strong constraints on the surface rotation as a function of time. This enables us to quantify the expected dispersion in lithium abundance and greatly reduces the sensitivity of the model predictions to uncertainties in the input physics. Those models which accurately reproduce the lithium abundances observed in open clusters can then be used to predict lithium depletion in halo stars. The general trend is that these ``open cluster normalized'' rotation models for halo stars predict more lithium depletion than do the standard models but less depletion than that predicted by earlier studies of rotational mixing with a less sophisticated treatment of angular momentum evolution. Our goal in this paper is to constrain the amount of lithium depletion in \\popii stars using several observables: (1) an estimate of the halo star initial rotation rate distribution as derived from open clusters, (2) the absence of large dispersion in the observed lithium abundances of the \\popii ``plateau'' stars, (3) the \\6li abundance and/or the \\6li/\\7li abundance ratio in HD 84937. We argue that, individually and in combination, these observables point to \\popii halo star lithium depletion of at least 0.2 dex (following from the open cluster data) but no more than 0.4 dex (a consequence of the narrowness of the plateau and of the \\6li considerations). Based on the observational data we can then infer the primordial lithium abundance and compare and contrast it with that predicted by standard BBN for consistency with the inferred primordial abundances of D and/or \\4he. \\section {Method and Comparison with Previous Models} \\subsection { Angular Momentum Evolution and Rotational Mixing} In standard stellar models, lithium depletion is a strong function of mass and composition; it also depends on the input physics, particularly the opacities and model atmospheres used to specify the surface boundary condition. In our models the standard model physics is the same as described in KPBS (\\cite{KPBS}): namely, interior opacities from OPAL (\\cite{RI92}), low-T opacities and model atmospheres at solar abundance from \\cite{Ku91a},\\cite{Ku91b}); nuclear reaction rates from \\cite{BPW95}, including the \\cite{CF88} \\7li(p, $\\alpha$) cross-section; Yale EOS; a solar calibrated mixing length, and $Y=0.235$. Chaboyer \\& Demarque (1994) noted that unusual Li-T$_{\\rm eff}$ trends for cool metal-poor stars occurred in models using the Kurucz atmospheres and could be lessened by using a grey atmosphere. We therefore ran halo star models with grey atmospheres and a mixing length of 1.25 as obtained for solar models constructed under the same assumptions. Stellar models which include rotational mixing require additional input physics beyond standard stellar models. The important new ingredients include: \\begin{enumerate} \\item a distribution of initial angular momenta \\item a prescription for angular momentum loss \\item a prescription for the internal transport of angular momentum and the associated mixing in radiative regions \\item the impact of rotation on the structure of the model. \\end{enumerate} There have been important changes in the treatment of the first three ingredients since the study of rotational mixing in halo dwarfs by \\cite{PDD}; in this section we discuss the ingredients of the current set of models and compare them with previous work. The treatment of angular momentum evolution in this set of models is the same as KPBS. The structural effects of rotation have been computed using the method of \\cite{KT70}, and are small for low mass stars which experience angular momentum loss. \\subsubsection {Initial Angular Momentum} The studies of lithium depletion in the presence of rotationally induced mixing by \\cite{PKSD} (the Sun), PKD (open cluster stars), and PDD (halo stars) all used a range of initial rotation rates in the pre-MS from the measured rotation velocities of young T Tauri stars (10-60 km/s) at a typical reference age of 1 Myr. No interaction between accretion disks and the protostar was included; the entire range of initial conditions was generated from the range in the initial rotation velocity. This caused some difficulty in reproducing the observed range in rotation (a factor of 20 in young open clusters), especially since the higher angular momentum loss rate in rapid rotators acted to reduce the range in initial rotation rates. This approach was also used by Chaboyer \\etal (1995a,b) for the Sun and open cluster stars respectively and by Chaboyer \\& Demarque (1994) for halo stars, although these papers used a different angular momentum loss law than the earlier work. As protostars evolve their moment of inertia decreases; this would produce an increase in rotation rate with decreased luminosity. In contrast, the rotation periods of classical T Tauri stars appear to be nearly uniform - they do not scale with luminosity as they would if the stars were conserving angular momentum as they contracted towards the main sequence (\\cite{B93}, \\cite{E93}). Weak-lined T Tauri stars, which are thought to be protostars without significant accretion disks, have much shorter rotation periods. This behavior would be expected if there were an accretion disk regulating the rotation of the central object and spinup only occurred when the disk is no longer magnetically coupled to the protostar (\\cite{K91}; \\cite{CCQ95}; \\cite{KMC95}). In this revised picture there is a constant surface rotation rate while there is a sufficiently massive disk around the central object; stars which detach from their disks earlier experience a larger change in angular velocity as they contract to the main sequence than those which detach from their disk later (and are only free to begin to spin up when their moment of inertia is smaller). This disk-locking hypothesis leads to a larger range in rotation rates for models of young open cluster stars, in better agreement with the data, and explains the very slow rotation of the majority of young stars (\\cite{BFA97}; \\cite{CCQ95}; \\cite{KMC95}; KPBS). These differences in the initial conditions lead to interesting consequences for the lithium depletion pattern. Although the initial rotation periods are comparable to those used in the earlier studies, the inclusion of star-disk coupling implies that the MS angular momentum is much lower for models with long disk lifetimes than it would be if the disks were absent. Disk lifetimes of 3 Myr are needed to match the rotation rates of the slow rotators in young clusters. The mean depletion for the majority of stars is therefore significantly lower than in previous studies because most young stars are slow rotators and the degree of mixing decreases for models with lower MS angular momentum. \\subsubsection {Angular momentum loss.} PKSD, PKD, and PDD all used a loss law of the form ${dJ}/{dt} \\propto \\omega^3$, where $\\omega$ is the angular velocity (\\cite{K88}). This implies greater angular momentum loss for rapid rotators than for slow rotators. However, models of this type do not reproduce the rapid rotators seen in young open clusters; the spin down of high angular momentum objects is too severe (PKD). On both theoretical and observational grounds, a more realistic functional form for the loss law is \\begin {equation} \\frac{dJ}{dt} \\propto \\omega^3\\ \\ \\ \\ \\ \\ (\\omega < \\omega_{\\rm crit}), \\end {equation} and \\begin {equation} \\frac{dJ}{dt} \\propto \\omega \\omega_{\\rm crit}^2 \\ \\ \\ \\ \\ \\ (\\omega > \\omega_{\\rm crit}) \\end {equation} (\\cite{MB91}; \\cite{CCQ95}; \\cite{M84}; \\cite{BS96}). Here $\\omega_{\\rm crit}$ is the angular velocity at which the angular momentum loss rate saturates; $\\omega_{\\rm crit}$ can be estimated from several observable quantities, thought to be correlated with surface magnetic field strength, which saturate at 5-20 times the solar rotation period (see \\cite{PS96}, \\cite{Kr97b} for reviews and discussions of recent work). A loss law which saturates at high rotation rates allows rapid rotation to survive into the early main sequence phase; a loss law of this form was adopted by Chaboyer \\etal (1995a,b) and Chaboyer \\& Demarque (1994). The observational data indicate that the duration of the rapid rotator phase is a function of mass in the sense that rapid rotation survives longer in lower mass stars. There is also indirect observational evidence for a mass-dependent saturation threshold (\\cite{PS96}; \\cite{Kr97b}). Models where the saturation threshold scales with the convective overturn time scale can reproduce the observed mass dependent spin down pattern. We adopt the mass-dependent saturation threshold of KPBS, \\begin {equation} \\omega_{\\rm crit} = \\omega_{\\rm crit}(\\odot) \\frac {\\tau_{\\rm conv}(\\odot)} {\\tau_{\\rm conv} (*)}, \\end {equation} where the convective overturn time scale $\\tau_{\\rm conv}$ as a function of ZAMS T$_{\\rm eff}$ was taken from the 200 Myr isochrones of Kim \\& Demarque (1996; KD). \\cite{KD96} only considered models at solar abundance, and we would need to recalibrate the KPBS models if we used a time and abundance dependent overturn timescale for the models. A comparison of the convection zone depth (as a function of T$_{\\rm eff}$, Z, and age) in the tables of PKD and PDD indicates that at young ages the convection zone depth at a given T$_{\\rm eff}$, and therefore the overturn time scale, depends only weakly on metal abundance. The value of $\\omega_{\\rm crit}$ is important primarily in the early main sequence. We therefore evolved standard halo star models to an age of 200 Myr and used the same $\\tau_{\\rm conv}$ as a function of T$_{\\rm eff}$ for the metal-poor models as reported by KD for the solar abundance models. We will consider models with a mass, composition, and time dependent overturn time scale in a future paper in preparation (\\cite{N98}). A loss law with a mass-dependent saturation threshold acts preferentially to suppress rapid rotation in the more massive stars, unlike the constant saturation threshold adopted by Chaboyer \\& Demarque (1994). This reduces both the absolute depletion and the dispersion of lithium abundances in hotter halo plateau stars relative to cooler plateau stars. \\subsubsection {Angular momentum transport.} The treatment of angular momentum transport is the same as in KPBS; a detailed review of the time scale estimates can be found in Pinsonneault (1997). We consider internal angular momentum transport by hydrodynamic mechanisms alone, and do not include potentially important mechanisms such as gravity waves (\\cite{KQ97}; \\cite{ZTM97}) and magnetic fields (\\cite{CM92}, \\cite{CM93}; \\cite{BCM98}). In previous work the degree of mixing was found to be insensitive to the assumptions about internal angular momentum transport (PKD; PDD; \\cite{CDP95a},b). \\cite{CVZ92} and \\cite{R96} obtained similar depletion patterns in models with solid body rotation, and \\cite{BCM98} reported similar results in models with magnetic fields and hydrodynamic mechanisms. This can be understood because the diffusion coefficients for angular momentum transport are actually largely determined by the balance between the surface boundary condition (the applied torque) and the flux of angular momentum from below a given shell. Theory gives an estimate of the ratio of the diffusion coefficients for mixing to those for angular momentum, which is related to the existence of anisotropic turbulence in stellar interiors (\\cite{CZ92}). The diffusion coefficients for mixing are calibrated on the Sun, but our lack of information on the rotational history of the Sun requires an exploration of models with varying solar initial conditions. ", "conclusions": "The lithium abundances of stars have interesting implications for cosmology and for the theory of stellar structure and evolution. The conclusions of this paper impact both of these areas. Standard stellar models implicitly assume that a variety of physical processes known to occur in real stars can be neglected; for many purposes this is a good approximation. There is now extensive evidence in \\popi stars that the predictions of standard models are not in agreement with all the data. Furthermore, work from a variety of investigators on physical effects not included in the standard models has concluded that there are physically well-motivated processes which can affect the surface lithium abundances of stars. Standard models and the previous generation of nonstandard models predicted initial \\7li abundances that differ by up to a factor of ten from the same set of current \\popii data, with very different implications for cosmology. We believe that mild envelope mixing driven by rotation is the most promising candidate for explaining the {\\it complete} observational picture. An improved treatment of angular momentum evolution in low mass stars now allows us to generate stellar models with rotation which are consistent with the rotation rates observed as a function of mass and age. We have inferred the distribution of initial angular momenta in young clusters and computed the distribution of lithium depletion factors as a function of mass, composition, and age. The resulting lithium depletion pattern is in good agreement with both \\popi and \\popii data, and the distribution of theoretical depletion factors is consistent with the distribution of abundances. For \\popii stars in particular, the observed slope of the [Li]-T$_{\\rm eff}$ relationship, the observed dispersion in abundance at fixed T$_{\\rm eff}$, the existence of a small population of overdepleted stars, and the simultaneous detection of \\6li and \\7li in one halo star are all consistent with the predictions of the theoretical models. The primary uncertainty in the absolute depletion of \\7li is the initial angular momentum of the Sun, which is used to calibrate the mixing diffusion coefficients. We have used the observed properties mentioned above to set bounds on \\7li depletion for \\popii stars. Those models which fit the \\popi data best predict a small but real depletion of lithium in the warm (T$_{\\rm eff} \\geq 5800$K), metal-poor ([Fe/H] $\\leq -1.3$) \\popii stars: 0.2 $\\leq$ logD$_{7} \\leq$ 0.4. These same models are consistent with the very small observed dispersion in abundances about the plateau value and with the survival of some \\6li. These depletion factors are significantly less than in previous models with rotational mixing (PDD, Chaboyer \\& Demarque 1994), and this difference can be directly attributed to the adoption of an improved treatment of angular momentum evolution. The uncertainties in the \\7li depletion factor can be reduced by a combination of new data and improved stellar evolution models. A large set of accurate lithium abundances in old open clusters would enable us to calibrate the diffusion coefficients for mixing without relying on the solar calibration. Microscopic diffusion and rotational mixing should be considered together, although the work of Chaboyer \\& Demarque (1994) indicated that the predictions of such models are similar to those which include rotation alone. The observed \\7li depletion pattern is relatively insensitive to the treatment of internal angular momentum transport, with a factor of ten in the time scale corresponding to less than a 10\\% change in the logarithmic lithium depletion factor in the work of PDD. However, the solar rotation curve does provide evidence for angular momentum transport from mechanisms not considered in the current paper, such as magnetic fields and/or internal gravity waves. Lithium depletion in models which include these effects and hydrodynamic mechanisms should be explored. Intermediate metallicity stars could also provide a valuable test of the dependence of lithium depletion on metal abundance. When our constraints on lithium-depletion are combined with the abundances inferred from observations of the Spite plateau halo stars ([Li]$_{\\rm OBS} = 2.25 \\pm$ 0.10) we find for the range in the abundance of primordial lithium: 2.35 $\\leq$ [Li]$_{\\rm P} \\leq$ 2.75 (2.2 $\\leq 10^{10}$(Li/H)$_{\\rm P} \\leq$ 5.6). Although lithium is not the ideal baryometer, a comparison with the predictions of standard BBN identifies two options: low-$\\eta$ and high-$\\eta$. The low-$\\eta$ branch suggests a very low baryon density, open Universe which may be in conflict with the baryon density inferred from observations of the Lyman-alpha forest. Although the helium abundance predicted for this low-$\\eta$ option agrees with that inferred from \\hii region data, the very high predicted deuterium abundance may not be consistent with observations. In contrast, the high-$\\eta$ branch corresponds to a baryon density consistent with the Ly-$\\alpha$ data and permits a higher density Universe. In this case the predicted primordial deuterium abundance is in excellent agreement with the low deuterium QSO data but a considerably higher primordial helium abundance is predicted than is inferred from some of the observational data." }, "9803/astro-ph9803245_arXiv.txt": { "abstract": "The low-redshift evolution of the intergalactic medium is investigated using hydrodynamic cosmological simulations. The assumed cosmological model is a critical density cold dark matter universe. The imposed uniform background of ionizing radiation has the amplitude, shape and redshift evolution as computed from the observed quasar luminosity function by Haardt \\& Madau. We have analysed simulated \\lya spectra using Voigt-profile fitting, mimicking the procedure with which quasar spectra are analysed. Our simulations reproduce the observed evolution of the number of \\lya absorption lines over the whole observed interval of $z=0.5$ to $z=4$. In particular, our simulations show that the decrease in the rate of evolution of \\lya absorption lines at $z\\le 2$, as observed by {\\em Hubble Space Telescope}, can be explained by the steep decline in the photo-ionizing background resulting from the rapid decline in quasar numbers at low redshift. ", "introduction": "Neutral hydrogen in the intergalactic medium produces a forest of \\lya absorption lines blueward of the \\lya emission line in quasar spectra (Lynds 1971). Observations of quasars at redshifts $z>2$ show that there is strong cosmological evolution in the number of \\lya lines, which can be characterised by a power law $dN/dz \\propto (1+z)^\\gamma$ (Sargent \\etal 1980), where $N$ is the number of lines above a threshold rest-frame equivalent width $W$ (typically $W>0.32\\AA$). Studies at high resolution using the Keck telescope find $\\gamma=2.78\\pm 0.71$ for $24$ for $z>4$ using the CTIO 4m telescope. In contrast at low redshifts, observations using the {\\em Hubble Space Telescope} find much less evolution, $\\gamma=0.48\\pm0.62$ for $z<1$ (Morris \\etal 1991, Bahcall \\etal 1991, 1993, Impey \\etal 1996). Recently, hydrodynamic simulations of hierarchical structure formation in a cold dark matter (CDM) dominated universe have been shown to be remarkably successful in reproducing this \\lya forest in the redshift range $z=4\\rightarrow 2$ (Cen \\etal 1994, Zhang, Anninos \\& Norman 1995, Miralda-Escud\\'e \\etal 1996, Hernquist \\etal 1996, Wadsley \\& Bond 1996, Zhang \\etal 1997, Theuns \\etal 1998). These simulations show that the weaker \\lya lines (neutral hydrogen column density $N_\\H\\le 10^{14}$ cm$^{-2}$) are predominantly produced in the filamentary and sheet-like structures that form naturally in this structure formation scenario. Velocity structure in these lines is often due to residual Hubble flow since many of the absorbing structures are expanding. In contrast, the stronger lines ($N_\\H\\ge 10^{16}$ cm$^{-2}$) tend to occur when the line of sight passes near a dense virialised halo. Over the redshift range investigated in these simulations a photo-ionizing background close to that inferred from quasars (Haardt \\& Madau 1996) is required to explain the properties of the \\lya forest. In fact, although most simulations have assumed a critical density, scale-invariant CDM universe, other variants of the CDM model provide acceptable fits with relatively small changes to the ionizing background (Cen \\etal 1994, Miralda-Escud\\'e \\etal 1996). The general success of CDM-like models in explaining the high redshift ($z\\ge 2$) properties of the \\lya forest is impressive. In this {\\em Letter} we investigate using hydrodynamic simulations whether a CDM universe with a photo-ionizing background dominated by quasars can explain the observed transition in the cosmological evolution of the number of \\lya lines at $z\\le 2$. ", "conclusions": "\\begin{figure*} \\setlength{\\unitlength}{1cm} \\centering \\begin{picture}(17,12) \\put(1, -4){\\special{psfile=fig2.ps hscale=65 vscale=65}} \\end{picture} \\caption{Evolution of the number of lines with given range in column density with redshift from simulations compared against observed evolution. Column density cut $10^{13.1}$cm$^{-2}\\le N_\\H\\le 10^{14}$ cm$^{-2}$-- simulations: circles connected with dotted line; data: open triangles (Kim \\etal 1997). Column density cut $10^{13.77}$ cm$^{-2}\\le N_\\H\\le 10^{16}$ cm$^{-2}$-- simulations: circles connected with solid line; data: filled squares (Bahcall \\etal 1993), open squares (Impey \\etal 1996), filled triangles (Kim \\etal 1997), filled pentagon (Lu \\etal 1996). Column density cut $10^{14.5}$ cm$^{-2}\\le N_\\H\\le 10^{16}$ cm$^{-2}$ -- simulation: circles connected with long dashed line; data: long dashed line shows evolution from Williger \\etal (1994). Large filled circles are simulation results for the low resolution, large box size, run, open circles are for a higher resolution, smaller box size, run. Large open pentagon: re-analysis of simulation at $z=0.5$, but imposing the ionizing background appropriate to $z=2$. Data were taken from figure~2 in Kim \\etal (1997), except for the Wiliger \\etal (1994) data. } \\label{fig:fig2} \\end{figure*} We show in figure~\\ref{fig:spectra} examples of simulated spectra at $z=3$, 2, 1 \\& 0.5 for the $L=22.22$ Mpc lower resolution simulation. Fluctuations in neutral hydrogen density, caused by gas tracing dark matter potential wells, produce absorption features similar to those in observed spectra. At low redshifts, there are large regions of the spectrum with very low absorption. These regions are separated by prominent absorption features, most of which are just a single strong line. At higher redshifts, there is considerable absorption over most of the spectrum and many lines are blended. Many of the strong lines at $z=0.5$ can be traced back to higher redshifts. There is a clear decrease in the comoving number of lines with decreasing redshift. The evolution of the column density distribution with redshift is illustrated in figure~\\ref{fig:fig1}. There is clear evolution in the simulated column density distribution with redshift. The rate of change depends on column density, with higher column density lines undergoing stronger evolution leading to steepening of the distribution. At $z=2$, the column density distribution is $\\propto N_\\H^{-1.6}$ whereas this has steepened to $\\propto N_\\H^{-2.1}$ at $z=0.5$ (see figure~\\ref{fig:fig1}). The rate of evolution also depends on redshift, with considerably stronger evolution at higher redshifts. The number of weak \\lya lines ($\\le 10^{13.1}$ cm$^{-2}$) remains approximately constant. At redshifts 3 and 2, there is good agreement between the simulated column density distribution for our higher resolution simulation ($L=5.5$~Mpc) and the observed one. The drop in the number of lines with redshift can be quantified by counting the number of lines within a given column density range. The simulation results are compared to observations in figure~\\ref{fig:fig2}. The simulations reproduce well the number of lines at a given redshift for all three column density cuts. Crucially, they also match very well the observed number of lines at low redshift. Consequently, the hierarchical picture of galaxy formation in a critical density universe, coupled with the observed evolution in the quasar luminosity function, can explain the observed evolution of the number of \\lya lines over the entire observed redshift range. We have re-analysed the $z=0.5$ output time after increasing the imposed ionization flux from its $z=0.5$ value to the value appropriate to $z=2$. The number of lines with $10^{13.77}$ cm$^{-2}\\le N_\\H\\le 10^{16}$ cm$^{-2}$ is shown by the open pentagon in figure~\\ref{fig:fig2}. This point falls onto the extrapolation for the number density evolution for $z\\ge 2$. Consequently, the dominant reason for the higher number of lines at low $z$ compared to what would be expected by extrapolation from high $z$, is the decrease in ionizing flux from $z=2$ to $z=0$, itself a consequence of the evolution of the quasar luminosity function. An estimate of the reliability of these simulations can be obtained by comparing the two simulations run at different resolutions. The higher resolution simulation produces more lines at all column densities, but the difference between the two simulations is well within the error bars of the observational results. This gives us confidence that we can reliable predict the number of lines from these simulations. In summary: our numerical simulations show that the properties of the \\lya forest are in excellent agreement with what is expected in a cold dark matter universe with a photo-ionizing background dominated by quasar light. In particular, our simulations show that the observed decrease in the rate of evolution of \\lya absorption lines at $z\\le 2$ can be explained by the steep decline in the photo-ionizing background resulting from the rapid decline in quasar numbers at low redshift." }, "9803/astro-ph9803303_arXiv.txt": { "abstract": "We report on the March-April 1997 BeppoSAX observations of Aql X-1, the first to monitor the evolution of the spectral and time variability properties of a neutron star soft X--ray transient from the outburst decay to quiescence. We observed a fast X--ray flux decay, which brought the source luminosity from $\\sim 10^{36}$ to $\\sim 10^{33}\\ergs$ in less than 10 days. The X--ray spectrum showed a power law high energy tail with photon index $\\Gamma\\sim 2$ which hardened to $\\Gamma \\sim 1-1.5$ as the source reached quiescence. These observations, together with the detection by RossiXTE of a periodicity of a few milliseconds during an X--ray burst, likely indicate that the rapid flux decay is caused by the onset of the propeller effect arising from the very fast rotation of the neutron star magnetosphere. The X--ray luminosity and hard spectrum that characterise the quiescent emission can be consistently interpreted as shock emission by a turned-on rotation-powered pulsar. ", "introduction": "Soft X--Ray Transients (SXRTs), when in outburst, show properties similar to those of persistent Low Mass X--Ray Binaries containing a neutron star (LMXRBs; White et al. 1984; Tanaka \\& Shibaza\\-ki 1996; Campana et al. 1998). The large variations in the accretion rate that are characteristic of SXRTs allow the investigation of a variety of regimes for the neutron stars in these systems which are inaccessible to persistent LMXRBs. While it is clear that, when in outbursts, SXRTs are powered by accretion, the origin of the low luminosity X--ray emission that has been detected in the quiescent state of several SXRTs is still unclear. An interesting possibility is that a millisecond radio pulsar (MSP) turns on in the quiescent state of SXRTs (Stella et al. 1994). This would provide a ``missing link\" between persistent LMXRBs and recycled MSPs. \\begin{figure*}[!htb] \\psfig{figure=aql_fig.ps,width=4.3cm} \\caption{ Light curve of the Feb.-Mar. 1997 outburst of Aql X-1 (panel {\\it a}. Data before and after MJD 50514 were collected with the RossiXTE ASM (2--10 keV) and the BeppoSAX MECS (1.5--10 keV), respectively. RossiXTE ASM count rates are converted to (unabsorbed) luminosities using a conversion factor of $4\\times 10^{35}\\ergs$ (before MJD 50512) and $2\\times10^{35}\\ergs$ (after MJD 50512) as derived from RossiXTE spectral fits (Zhang, Yu \\& Zhang 1998). BeppoSAX luminosities are derived directly from the spectral data (see text). The evolution of the flux from MJD 50480 to MJD 50512 is well fit by a Gaussian centered on MJD 50483.2. This fit however does not provide an acceptable description for later times (see the dot-dashed line), not even if the accretion luminosity is calculated in the propeller regime (dashed line). The straight solid line represents the X--ray luminosity corresponding to the closure of the centrifugal barrier $L_{\\rm min}$ (for a magnetic field of $10^8$ G and a spin period of 1.8 ms) and the straight dashed line the luminosity gap due to the action of the centrifugal barrier, $L_{\\rm cor}$. The dotted line marks the minimum luminosity in the propeller regime ($L_{\\rm lc}$). Panel {\\it b} shows the BeppoSAX unfolded spectra of Aql X-1 during the early stages of the fast decline (1) and during the quiescent phase (3--6, summed). The best fit spectral model (black body plus power law) is superposed to the data.} \\label{tot} \\end{figure*} Aql X-1 is the most active SXRT known: more than 30 X--ray and/or optical outbursts have been detected so far. The companion star has been identified with the K1V variable star V1333 Aql and an orbital period of 19 hr has been measured (Chevalier and Ilovaisky 1991). The outbursts of Aql X-1 are generally characterised by a fast rise (5--10 d) followed by a slow decay, with an $e-$folding time of 30--70~d (see Tanaka \\& Shibazaki 1996 and Campana et al. 1998 and references therein). Type I X--ray bursts were discovered during the declining phase of an outburst (Koyama et al. 1981), testifying to the presence of a neutron star. Peak X--ray luminosities are in the $\\sim (1-4)\\times 10^{37}\\ergs$ range (for the $\\sim 2.5$ kpc distance inferred from its optical counterpart; Thorstensen et al. 1978). Close to the outburst maximum the X--ray spectrum is soft with an equivalent bremsstrahlung temperature of $k\\,T_{\\rm br} \\sim 4-5$~keV. Sporadic observations of Aql X-1 during the early part of the outburst decay (Czerny et al. 1987; Tanaka \\& Shibazaki 1996; Verbunt et al. 1994) showed that when the source luminosity drops below $\\sim 10^{36}\\ergs$ the spectrum changes to a power law with a photon index of $\\Gamma\\sim 2$, extending up to energies of $\\sim 100$ keV (Harmon et al. 1996). ROSAT PSPC observations revealed Aql X-1 in quiescence on three occasions at a level of $\\sim 10^{33}\\ergs$ (0.4--2.4 keV; Verbunt et al. 1994). In this lower energy band the spectrum is considerably softer and consistent with a black body temperature of $k\\,T_{\\rm bb} \\sim 0.3$~keV. ", "conclusions": "\\begin{table*} \\begin{center} \\caption{Summary of SAX NFIs observations.} \\label{tab1} \\begin{tabular}{cccccc} Obs./Date & LECS/MECS-PDS & LECS Count Rate & MECS Count Rate & PDS Count Rate\\\\ & Expos. (s) & (c s$^{-1}$) & (c s$^{-1}$) & (c s$^{-1}$)\\\\ \\hline 1/March 8$^{th}$, 1997 &\\, 5240/21342 & $0.84 \\pm 0.02$ & $2.2\\pm 0.01$ & $0.87\\pm0.06$ \\\\ 2/March 12$^{th}$, 1997 &\\, 3247/21225 & $(2.5\\pm0.4) \\times 10^{-2}$ & $(7.4\\pm 0.2) \\times 10^{-2}$ & $\\lsim 0.19$ \\\\ 3/March 17$^{th}$, 1997 &\\, 5755/17258 & $(5.6\\pm1.7) \\times 10^{-3}$ & $(1.3\\pm 0.1) \\times 10^{-2}$ & $\\lsim 0.24$ \\\\ 4/March 20$^{th}$, 1997 &\\, 4287/22589 & $(5.8\\pm1.9) \\times 10^{-3}$ & $(1.6\\pm 0.1) \\times 10^{-2}$ & $\\lsim 0.17$\\\\ 5/April 2$^{nd}$, 1997 &\\, 8440/23576 & $(6.2\\pm1.3) \\times 10^{-3}$ & $(1.2\\pm 0.1) \\times 10^{-2}$& $\\lsim 0.17$ \\\\ 6$^{\\rm o}$/May 6$^{th}$, 1997 & 11789/21703 & $(6.7\\pm1.1) \\times 10^{-3}$ & --- & $\\lsim 0.19$ \\\\ \\hline \\end{tabular} \\end{center} {\\noindent $^{\\rm o}$ No MECS data were obtained.} \\end{table*} \\begin{table*} \\begin{center} \\caption{Summary of spectral fits.} \\label{tab2} \\begin{tabular}{ccccccc} Obs.$^*$ & Black body & Black body & Power law & PL/BB & Mean Luminosity$^\\dag$ & Reduced\\\\ & $k\\,T_{\\rm bb}$ (keV) & $R_{\\rm bb}$ (km)& $\\Gamma$ & flux ratio$^{\\ddag}$&(erg s$^{-1}$)& $\\chi^2$\\\\ \\hline 1 & $0.42\\pm0.02$&$2.6\\pm0.3$&$1.9\\pm0.1$& 3.7 & $9\\times 10^{34}$ &1.0 \\\\ 2 & $0.3_{-0.2}^{+0.1}$&$0.7_{-0.3}^{+6.6}$&$1.8\\pm0.7$& 1.6 & $2\\times 10^{33}$ &0.9 \\\\ \\ 3--6 & $0.3\\pm0.1$&$0.8_{-0.1}^{+0.4}$&$1.0\\pm0.3$& 0.7 & $6\\times 10^{32}$ &1.3 \\\\ \\hline \\end{tabular} \\end{center} { \\noindent Errors are $1\\,\\sigma$. \\noindent $^*$ Spectra from the LECS and MECS (and PDS for the first observation) detectors have been considered. The spectra corresponding to the quiescent state have been summed up, in order to increase the statistics and an upper limit from the summed PDS data was also used to better constrain the power law slope. \\noindent $\\dag$ Unabsorbed X--ray luminosity in the energy range 0.5--10 keV. In the case of the first observation the PDS data were included in the fit (the unabsorbed 0.5--100 keV luminosity amounts to $2\\times 10^{35}\\ergs$). \\noindent $\\ddag$ Power law to black body flux ratio in the 0.5--10 keV energy range.} \\end{table*} The BeppoSAX observations enabled us to follow for the first time the evolution of a SXRT outburst down to quiescence. The sharp flux decay leading to the quiescent state of Aql X-1 is reminiscent of the final evolution of dwarf novae outbursts (e.g. Ponman et al. 1995; Osaki 1996), although there are obvious differences with respect to the X--ray luminosities and spectra involved in the two cases, likely resulting from the different efficiencies in the gravitational energy release between white dwarfs and neutron stars. Models of low mass X--ray transient outbursts hosting an old neutron star or a black hole are largely built in analogy with dwarf novae outbursts. In particular, van Paradijs (1996) showed that the different range of time-averaged mass accretion rates over which the dwarf nova and low mass X--ray transient outbursts were observed to take place is well explained by the higher level of disk irradiation caused by the higher accretion efficiency of neutron stars and black holes. However, the outburst evolution of low mass X--ray transients presents important differences. In particular, the steepening in the X--ray flux decrease of Aql X-1 has no clear parallel in low mass X--ray transients containing Black Hole Candidates (BHCs). The best sampled light curves of these sources show an exponential-like decay (sometimes with a superposed secondary outburst) with an $e-$folding time of $\\sim 30$ d and extending up to four decades in flux, with no indication of a sudden steepening (Chen et al. 1997). In addition, BHC transients display a larger luminosity range between outburst peak and quiescence than neutron star SXRTs (Garcia et al. 1998 and references therein). Being the mass donor stars and the binary parameters quite similar in the two cases, it appears natural to attribute these differences to the different nature of the underlying object: neutron stars possess a surface and, likely, a magnetosphere, while BHCs do not. When in outburst accretion down to the neutron star surface takes place in SXRTs, as testified by the similarity of their properties with those of persistent LMXRBs, especially the occurrence of type I bursts and the X--ray spectra and luminosities. The mass inflow rate during the outburst decay decreases, causing the expansion of the magnetospheric radius, $r_{\\rm m}$. Thus, accretion onto the neutron star surface can continue as long as the centrifugal drag exerted by the corotating magnetosphere on the accreting material is weaker than gravity (Illarionov \\& Sunyaev 1975; Stella et al. 1986). This occurs above a limiting luminosity $L_{\\rm min}=G\\,M\\,\\mdot_{\\rm min}/R \\sim 4\\times10^{36} \\,B_8^2\\,P_{-3}^{-7/3} ~{\\rm erg\\,s^{-1}}$, where $G$ is the gravitational constant; $M$, $R$, $B=B_8\\,10^8$ G and $P=P_{-3}\\,10^{-3}$ ms are the neutron star mass, radius, magnetic field and spin period, respectively (here and in the following we assume $M=1.4\\msole$ and $R=10^6$ cm). As $r_{\\rm m}$ reaches the corotation radius, $r_{\\rm cor}$, accretion onto the surface is inhibited and a lower accretion luminosity ($} 40$ kpc from the Galactic center, with a diameter ${>} 20$ kpc and a mass ${>} 10^8$ solar masses. We then discuss a number of other clouds that are positionally associated with the Local Group galaxies. The kinematics of the entire ensemble of HVCs is inconsistent with a Galactic origin, but implies that the HVCs are falling towards the Local Group barycenter. The HVCs obey an angular--size/velocity relation consistent with the Local Group infall model. We simulate the dynamical evolution of the Local Group. The simulated properties of material falling into the Local Group reproduce the location of two of the three most significant groupings of clouds and the kinematics of the entire cloud ensemble (excluding the Magellanic Stream). We interpret the third grouping (the A, C, and M complexes) as tidally unstable nearby material falling onto the Galactic disk. We interpret the more distant HVCs as dark matter ``mini--halos'' moving along filaments towards the Local Group. Most poor galaxy groups should contain HI structures to large distances bound to the group. We suggest that the HVCs are local analogues of the Lyman--limit absorbing clouds observed against distant quasars. We argue that there is a Galactic fountain in the Milky Way, but that the fountain does not explain the origin of the HVCs. Our analysis of the HI data leads to the detection of a vertical infall of low--velocity gas towards the plane. We suggest that the fountain is a local phenomenon involving neutral gas with characteristic velocities of 6 \\kmse rather than 100 \\kms. This implies that the chemical evolution of the Galactic disk is governed by episodic infall of metal-poor HVC gas that only slowly mixes with the rest of the interstellar medium. The Local--Group infall hypothesis makes a number of testable predictions. The HVCs should have sub-solar metallicities. Their H$\\alpha $ emission should be less than that seen from the Magellanic Stream. The clouds should not be seen in absorption to nearby stars. The clouds should be detectable in both emission and absorption around other groups. We show that current observations are consistent with these predictions and discuss future tests. ", "introduction": "\\label{sec:intro} Since their discovery in 1963 by \\markcite{Muller63}Muller, Oort, \\& Raimond, the nature of the high--velocity hydrogen clouds (HVCs) has remained a mystery. HVCs are clouds that deviate from Galactic circular rotation by as much as several hundred kilometers per second. Although astronomers have speculated about the origin of HVCs since their detection, no single explanation has encompassed the vast quantity of data that has been collected on the clouds (see \\markcite{WvW97} Wakker \\& van Woerden 1997, hereafter WvW97, and references therein). Particularly important is the lack of agreement on a characteristic distance for the clouds, because most of the relevant physical parameters depend on distance to one order or another. In the 1970s, a well--defined subset of the clouds was identified as a tidal stream associated with the Magellanic Clouds (\\markcite{Mathewson74}Mathewson, Cleary, \\& Murray 1974), but since then no consensus has arisen regarding the nature of the remaining clouds which constitute the majority of HVCs. In this paper, we suggest that the HVCs represent infall of the intergalactic medium onto the Local Group. Previous authors have explored the possibility that the HVCs are infalling primordial gas (\\markcite{Oort66,Oort70}Oort 1966, 1970) and have associated the HVCs with the Local Group (\\markcite{Verschuur69}Verschuur 1969; \\markcite{Kerr69}Kerr \\& Sullivan 1969; \\markcite{Arp85}Arp 1985; \\markcite{Bajaja87}Bajaja, Morras, \\& P\\\"oppel 1987; \\markcite{Arp91}Arp \\& Sulentic 1991), but subsequent observations always produced fundamental difficulties for the particular models considered. Here, we assemble evidence based on new general--purpose surveys of atomic hydrogen gas by \\markcite{Stark92}Stark et al. (1992) and by \\markcite{Hartmann97}Hartmann \\& Burton (1997), and on the HVC surveys by \\markcite{Hulsbosch88}Hulsbosch \\& Wakker (1988) and by \\markcite{Bajaja85}Bajaja et al. (1985), and consider theoretical arguments in the context of modern cosmology. We argue that the clouds are matter accreting onto the Local Group of galaxies. Their velocities would thus largely reflect the motion of the clouds in the gravitational potential of the Local Group and the motion of the LSR about the Galactic center. We suggest that the clouds represent the building blocks from which the Local Group was assembled and that they continue to fuel star formation in the disk of the Milky Way. The evidence is presented as follows. In Section 2, we review some of the observed properties of the high--velocity clouds, and indicate those which are not consistent with a Galactic origin for the HVCs. In Section 3, we detail the observations entering our analysis. In Section 4, we discuss the stability of the HVCs against gravitational collapse and against Galactic tidal forces, and suggest that these considerations imply that the most of the clouds are extragalactic at distances typical of the Local Group. The stability criteria imply, however, that the largest clouds are nearby and possibly are interacting with the Milky Way. In Section 5, we discuss three individual clouds, one of which must be beyond the disk of the Milky Way, and two others that appear to be associated with M31 and M33, suggesting that at least some of the HVCs may be extragalactic. We identify a subset of the HVCs centered near the barycenter of the Local Group, and show that its kinematics as well as that of the entire HVC ensemble are well described as being at rest with respect to the Local--Group Standard of Rest (LGSR); the kinematics are thus inconsistent with a Galactic origin. The entire HVC ensemble is also shown to exhibit an angular--size/velocity relation consistent with membership in the Local Group. In Section 6, we simulate the accretion history of the Local Group and show that the simulation accounts for the observed distribution and kinematics of the HVC ensemble. The agreement between the simulation and the observations supports inferences about similarities between the Local Group HVCs and the Ly--$\\alpha$ absorbing clouds observed toward quasars. We show that the velocity extrema observed for the HVCs are consistent with their membership in the Local Group. In Section 7, we discuss the distances and abundances of the HVCs in the context of the Local Group HVC hypothesis, and show that the hypothesis is consistent with all of the observations made to date. We review extragalactic HI searches for HVCs which have revealed clouds with properties similar to those we derive, in about the expected numbers. In Section 8, we discuss the implied mass accretion rate, and implications for the chemical evolution of the disk of the the Milky Way. We also present evidence for the Galactic fountain in low--velocity HI which suggests that the HI disk of the Milky Way is not in hydrostatic equilibrium. In Section 9, we conclude by discussing predictions and future tests of the model, and summarize the principal arguments made in this paper. ", "conclusions": "\\label{sec:finale} Most cosmologists believe that galaxy formation is a hierarchical process: galaxies grow by accreting small clouds of gas and dark matter. This process is a continuing one and we expect that galaxies and groups are currently accreting new clouds. We simulated this process for the Local Group and found that properties of the accreted clouds are similar to certain properties of the high--velocity--cloud phenomenon (excluding the Magellanic Stream HVCs): $\\bullet $ Most of the HVCs are located either near the general direction of M31, towards the barycenter of the Local Group, or in the antibarycenter direction, some 180\\dege from the direction of M31 (see Figure~\\ref{fig:lb}). $\\bullet $ HVCs have chemical abundances similar to that of intra--group gas, and different from the abundances characteristic of the inner Galaxy. If HVCs were ejected from the inner Galaxy as part of a Galactic fountain, then their metal abundance would exceed the solar value, and this is not observed. $\\bullet $ HVCs have an angular--size/velocity relation that is consistent with the clouds being nearly self--gravitating, and at a distance of $\\sim 300$ kpc. If the Local--Group HVC hypothesis discussed in this paper is correct, then studies of HVCs can directly probe the process of galaxy formation. The validity of this hypothesis can be tested by a number of future observations: $\\bullet $ Most observations of nearby galaxies would not have detected the gas clouds that are equivalent to the HVCs. Moreover, many HI maps of external galaxies extend just beyond the Holmberg radius. Our discussion would have the typical HVC located nearly a Mpc from the galactic center. It will be interesting to test our hypothesis with deep HI observations of isolated groups and filaments, searching for HI clouds associated with groups, rather than with individual galaxies. $\\bullet $ Lyman--limit clouds, which are seen in absorption towards distant quasars, have column densities similar to those of the HVCs. Observations of nearby Lyman--limit and Lyman--$\\alpha $ systems show that they are not all associated directly with individual galaxies, but rather with groups of galaxies (\\markcite{Oort81}Oort 1981; \\markcite{Stocke95}Stocke et al. 1995; \\markcite{vanGorkom96}van Gorkom et al. 1996; \\markcite{Rauch96}Rauch, Weymann, \\& Morris 1996). In the scenario outlined in this paper, we expect that these clouds would have properties similar to those of the local HVCs. Thus, it would be interesting to use STIS to look for lower--column--density high--velocity HI clouds, which would correspond to the Lyman--$\\alpha $ clouds. $\\bullet $ The simulations predict large amounts of gas accreting onto M31 and the Local Group from the region of space beyond M31, under the gravitational attraction of both M31 and our own Galaxy. Because this gas is several Mpc away, the gas clouds are expected to have small angular sizes and relatively low column densities. Deep HI observations in the M31 direction should be able to detect this gas. The hypothesis central to this paper, namely that HVCs are at distances of around 1 Mpc, would be falsified by the detection of absorbtion in an HVC seen against stars in the Milky Way halo in the direction of M31 or in the anti--M31 direction. On the other hand, further measurements of low levels of H$\\alpha$ emission towards these HVCs will strengthen the case for their extragalactic nature." }, "9803/astro-ph9803067_arXiv.txt": { "abstract": "A solitary millisecond pulsar, if near the mass limit, and undergoing a phase transition, either first or second order, provided the transition is to a substantially more compressible phase, will emit a blatantly obvious signal---spontaneous spin-up. Normally a pulsar spins down by angular momentum loss to radiation. The signal is trivial to detect and is estimated to be ``on'' for 1/50 of the spin-down era of millisecond pulsars. Presently about 25 solitary millisecond pulsars are known. The phenomenon is analogous to ``backbending'' observed in high spin nuclei in the 1970's. ", "introduction": "The formation of a new phase of matter, a softer one than nuclear matter, may cause a rapidly rotating pulsar to produce a prolonged signal that is dramatic, easy to detect and easy to understand \\cite{glen97:a}. The most plausible high density phase transition is deconfinement as predicted by QCD \\cite{asymptotic}. The signal I will describe will occur for either a first or second order transition so long as it is accompanied by a sufficient softening of the \\eosp. (Cf. Fig.\\ \\ref{eos}.) Strictly speaking we do not even know that quarks can be deconfined under extreme conditions or otherwise. It is an `expectation' based on the QCD property of asymptotic freedom \\cite{asymptotic}. We would like to prove that this phase is a possible phase of matter. If so, it would have pervaded the very early universe, but quark confinement in hadrons occurred at an early time and the thermal equilibrium that existed then leaves no signal today. \\begin{figure}[htb] \\vspace{-.25in} \\begin{minipage}[t]{80mm} \\makebox[79mm] {\\psfig{figure=ps.qm1,width=3in,height=3.6in} } \\caption {The \\eos (labeled `Hybrid') of neutron star matter with a first order deconfinement phase transition. The normal phase contains nucleons, hyperons and leptons in equilibrium. The mixed phase contains as well, the three light flavor quarks. Comparison is made with a case in which deconfinement is not taken account of (labeled `n+p+H'). (Nuclear properties include the observed binding, saturation density, symmetry energy and $K=300$ MeV, $m^\\star_{{\\rm sat}}/m=0.7$) \\label{eos_k300_y_h} \\label{eos} } \\end{minipage} \\hspace{\\fill} \\begin{minipage}[t]{75mm} \\makebox[74mm] {\\psfig{figure=ps.qm2,width=3in,height=3.6in}} \\caption { Density profiles of two stars of the same mass $M=1.42 \\msun$ but differing composition; (1) Hyperon star (neutron-proton-hyperon-lepton), (2) Hybrid (a pure quark matter core surrounded by mixed phase and outer pure hadronic confined phase). \\Eoss as in Fig.\\ \\protect\\ref{eos_k300_y_h}. Interior differences are dramatic but not directly measureable. (For a description of neutron star matter and relativistic stars see Ref.\\ \\protect\\cite{book}. \\label{prof_k240} } \\end{minipage} \\end{figure} From the balance of gravitational and centrifugal forces on a particle at the surface of rapidly rotating stars such as the millisecond pulsars, we know that the central density is a few times nuclear, the same range of energy densities as are expected to be produced in relativistic nuclear collisions. Let us assume that the critical deconfinement density occurs in the density range spanned by spherical stars and therefore in the population of slow pulsars to which the Crab belongs and of which there are about 800 presently known. In this case newly born neutron stars have a quark core essentially from birth. But we have no way to tell if this is indeed the case. Models of neutron stars having different composition generally differ in the range of masses and radii permitted, and their density profiles may be very different (cf. Fig.\\ \\ref{prof_k240}). However as far as measureable properties are concerned, ordinary neutron stars and hybrid stars (neutron stars with a quark core) are practically indistinguishable. Cooling rates are presumeably sensitive to the internal composition, but theoretical estimates are very uncertain. I will sketch the generally accepted evolutionary life of pulsars \\cite{heuvel91:a}. There are two distinct populations of pulsars (see Figs.\\ \\ref{period} and \\ref{pulsar}), the canonical pulsars (about 800 of them now known) with periods between 1/50 sec and 8 sec, and the millisecond pulsars (about 50), which are believed to be more evolved than the former. Stars are born into the first of these populations, and may, on a very long time-scale, evolve into the second. (See Fig.\\ \\ref{pulsar}.) \\begin{figure}[htb] \\vspace{-.25in} \\begin{center} \\begin{minipage}[t]{130mm}\\hspace{.8in} \\makebox[79mm] {\\psfig{figure=ps.qm3,width=4.2in,height=3.5in} } \\caption { Distribution of pulsar periods. The lower group consists of `recycled' millisecond pulsars, the higher to the canonical pulsars in the first stage of their evolution (see Fig.\\ \\protect\\ref{pulsar}). \\label{period} } \\end{minipage} \\end{center} \\vspace{-.35in} \\end{figure} As the stellar core of a luminous star collapses to form a neutron star, it is spun up by conservation of angular momentum and acquires an enormous magnetic field of $10^{12}{\\rm~to~}10^{13}$ gauss because of flux conservation. The star is born as a rotating magnetic dipole. It has a tremendous store of rotational energy that will keep it spinning for 10 million years. The electromagnetic radiation beamed along the spinning dipole is what we see as pulsed radio emission once each rotation, if as observers, we lie on the cone swept out by the beam. \\begin{figure}[htb] \\vspace{-.25in} \\begin{center} \\begin{minipage}[t]{130mm}\\hspace{.8in} \\makebox[79mm] {\\psfig{figure=ps.qm4,width=4.2in,height=3.5in} } \\vspace{-.35in} \\caption { Pulsars are born with field $B\\sim 10^{12} {\\rm~to~} 10^{13}$ gauss and evolve toward the right. Periods change with time as $\\sim 1/P$ and so pulsars accumulate at large $P$. They become radio silent in about $10^7$ years and remain stagnant until they capture a companion star, or unless they had one all along. Accretion from the less dense companion spins them up along a line like `silent evolution'. A combination of the now weaker field but higher frequency turns them on again as `recycled' millisecond pulsars. They then again evolve toward shorter period, but now on a very long time-scale because of the weaker field. \\label{pulsar} } \\end{minipage} \\end{center} \\vspace{-.35in} \\end{figure} In a plot of magnetic field $B$ vs period $P$ (Fig.\\ \\ref{pulsar}), stars move from top left to right because of loss of angular momentum to radiation. They disappear as active pulsars when a combination of angular velocity and field strength is insufficient to produce radiation. It takes about $10^7$ years to complete this first phase. However, either from birth, or afterward, the star may have or capture a lower density companion as is often evidenced by the presence of an orbiting white dwarf. During an accretion era, the compact star is spun up by infalling matter that it tears off from its less dense companion. It looses some magnetic field during accretion, perhaps by ohmic resistance during the long radio silent era. In this `silent' era the the star moves diagonally from top right to bottom left in the $B-P$ diagram. The neutron star becomes centrifugally flattened as it approaches millisecond periods. The central density falls. The core of quark matter shrinks as quarks recombine to form baryons. The quark core may disappear altogether. The silent neutron star, having completed a part of its life cycle, turns on again as a millisecond pulsar of low magnetic field when the lower field but higher angular velocity can once again produce radiation. Presently 50 such pulsars have been discovered, half of which still have a binary companion. The number of millisecond pulsars presently known is believed to be a fraction of the total population because of search selection effects. Millisecond pulsars, which have weaker fields ($10^8{\\rm~to~}10^9$ gauss), spin down very slowly since the deceleration is proportional to $B^2$. Their characteristic age is $P/2\\dot{P}\\sim 10^9$ years. The central density is initially centrifugally diluted but as it spins down, the central density will rise again and the critical density will be reached, first at the center, and then in an expanding region. The growth of the central region of deconfined matter is paced by the slow spin-down, slow because of the coupling of rotation of the stellar magnetic dipole to electromagnetic processes. Stiff nuclear matter is being replaced in the core by highly compressible quark matter. The weight of the overlaying layers of nuclear matter weigh down on the core and compress it. Its density rises. The star shrinks---mass is redistributed with growing concentration at the center. The by-now more massive central region gravitationally compresses the outer nuclear matter even further, amplifying the effect. The density profile for a star at three angular velocities, (1) the limiting Kepler velocity which is stretched in the equatorial plane and centrally diluted, (2) an intermediate angular velocity, and (3) a non-rotating star, are shown in Fig.\\ \\ref{prof_k300b180}. We see that the central density rises with decreasing angular velocity by a factor of three and the equatorial radius decreases by 30 percent. In contrast, for a model for which the phase transition did not take place, the central density would change by only a few percent \\cite{weber90:d}. The phase boundaries are shown in Fig.\\ \\ref{omega_r_k300B180} from the highest rotational frequency to zero rotation. \\begin{figure}[htb] \\vspace{-.5in} \\begin{minipage}[t]{80mm} \\makebox[79mm] {\\psfig{figure=ps.qm5,width=3in,height=3.6in} } \\caption { Mass profiles as a function of equatorial radius of a star rotating at three different frequencies, as marked. At low frequency the star is very dense in its core, having a 4 km central region of highly compressible pure quark matter. At intermediate frequency, the pure quark matter phase is absent and the central 8 km is occupied by the mixed phase. At higher frequency (nearer $\\Omega_K$) the star is relatively dilute in the center and centrifugally stretched. Inflections at $\\epsilon=220$ and $950$ are the boundaries of the mixed phase. \\label{prof_k300b180} } \\end{minipage} \\hspace{\\fill} \\begin{minipage}[t]{75mm} \\makebox[74mm] {\\psfig{figure=ps.qm6,width=3in,height=3.6in} } \\caption { Radial boundaries at various rotational frequencies separating (1) pure quark matter, (2) mixed phase, (3) pure hadronic phase, (4) ionic crust of neutron rich nuclei and surface of star. The pure quark phase appears only when the frequency is below $\\Omega \\sim 1370$ rad/s. Note the decreasing radius as the frequency falls. The frequencies of two pulsars, the Crab and PSR 1937+21 are marked for reference. \\label{omega_r_k300B180} } \\end{minipage} \\end{figure} The redistribution of mass and shrinkage of the star change its moment of inertia and hence the characteristics of its spin behavior. The star must spin up to conserve angular momentum which is being carried off only slowly by the weak electromagnetic dipole radiation. The star behaves like an ice skater who goes into a spin with arms outstretched, is slowly spun down by friction, temporarily spins up by pulling the arms inward, after which friction takes over again. It is that simple to describe, and that is the blatant signal I mentioned---the spontaneous spin-up of an isolated millisecond pulsar that is radiating angular momentum and ought otherwise to be slowing down. ", "conclusions": "An isolated millisecond pulsar will spin up over an epoch of $2 \\times10^7$ years out of a spin-down life of $10^9$ years if it undergoes a phase transition obeying the two conditions (i) the transition causes a substantial softening of the \\eosp, and (ii) the critical density is attained in stars very near the mass limit. The spin-up epoch, compared to the spin-down life of the pulsar, corresponds to an `event rate' of 1/50. The determination of whether a pulsar is spinning up or down is trivial. Of the presently known millisecond pulsars, about 25 are isolated. We are approaching the moment of truth for this observable signal of a phase transition. We have emphasized that the transition need not be of first order so long as it is accompanied by a sufficient softening of the \\eosp. We do not have a measure of what we mean by this. Of course our model does possess the requisite softening, or our results would not have exhibited backbending. If no pulsar is observed to produce the signal, little is learned. Just as in the search for deconfinement in high energy nuclear collisions, failure to observe a signal does not inform us that the deconfined phase does not exist." }, "9803/astro-ph9803192_arXiv.txt": { "abstract": "We construct evolutionary synthesis models for simple stellar populations using the evolutionary tracks from the Padova group (1993, 1994), theoretical colour calibrations from Lejeune et al. (\\cite{lejeune}, \\cite{lejeune1}) and fit functions for stellar atmospheric indices from Worthey et al. (\\cite{worthey}). A Monte-Carlo technique allows us to obtain a smooth time evolution of both broad band colours in UBVRIK and a series of stellar absorption features for Single Burst Stellar Populations ({\\bf SSPs}). We present colours and indices for SSPs with ages from $1 \\cdot 10^9$ yrs to $1.6 \\cdot 10^{10}$ yrs and metallicities $[M/H]$ = -2.3, -1.7, -0.7, -0.4, 0.0 and 0.4. Model colours and indices at an age of about a Hubble time are in good agreement with observed colours and indices of the Galactic and M 31 GCs. ", "introduction": "Colour distributions of Globular Cluster ({\\bf GC}) systems are observed for a large number of early-type galaxies (E, S0, dE, cD) using ground-based Washington or HST broad band photometry. In most cases double-peak or broad/mul\\-ti-peak colour distributions are seen (e.g. Zepf \\& Ashman \\cite{zepf}, Elson \\& Santiago \\cite{elson}, Kissler-Patig \\etal \\cite{kissler}). If the different colour subpopulations of GCs are formed in different events then they may contain clues to the formation history of their parent galaxies. For example a two-peak colour distribution may result, if in addition to a primary initial collapse population of GCs, a secondary population of GCs were formed either in a merger-induced starburst (Schweizer \\cite{schweizer}, Ashman \\& Zepf \\cite{ashman}, Fritze -- v. Alvensleben \\& Gerhard \\cite{fritze1}, Fritze -- v. Alvensleben \\& Burkert \\cite{fritze2}) or else in some distinct secondary phase of cluster formation within the original galaxy (Forbes \\etal \\cite{forbes}). Likewise the broad or multi-peaked colour distribution often observed in GC systems around cD galaxies may point to a series of GC formation events during the hierarchical assembly of the parent galaxy or to some protracted GC formation or accretion mechanism. A well-known difficulty with the interpretation of colour distributions is the degeneracy of colours with respect to age and metallicity. While for Washington photometry there are well established and reliable calibrations of colours in terms of metallicity, the situation with HST broad band observations of GC systems is less clear. A better understanding of the formation of composite GC systems would be possible if separate age and metallicity distributions could be disentangled from an observed colour distribution. A second issue concerns the interpretation of colours for young star cluster systems detected with HST in many interacting and starburst galaxies. The question is, if these YSC systems -- at least some fraction of them -- are the progenitors of GC systems. In an attempt to answer this question star clusters are being imaged with HST in an age sequence of interacting galaxies -- from early stages of interaction through merger remnants up to E/S0s (eg. Schweizer \\etal \\cite{schweizer1}, Whitmore \\etal \\cite{whitmore}, Miller \\etal \\cite{miller}). With 10 m class telescopes, spectroscopy of the brighter members of young star cluster populations is becoming possible (Kissler-Patig \\etal \\cite{kissler1}, Brodie \\etal \\cite{brodie}, but see also 4 -- 5 m class spectra by Schweizer \\& Seitzer \\cite{schweizer} or Zepf \\etal \\cite{zepf1}). Spectroscopy will only be possible for a subsample of YSCs. Thus the determination of ages and metallicities from broad band colors will still be necessary. It is thus desirable to study the evolution of broad band colours and absorption indices for single burst stellar populations of various metallicities using the most recent and complete stellar evolutionary tracks as well as careful colour and index calibrations.This allows one to obtain theoretical calibrations of broad band colours and indices in terms of metallicity over the full range of ages under investigation, i.e. from $10^7$ yr to a Hubble time. Since theoretical calculations for the evolution of stars are only available for a discrete grid of masses, some means for obtaining a smooth evolution of the composite population is needed. Applying the tracks as they are would create discontinuities because all stars of a given mass would reach the giant branch at the same time, dominating the integrated light until they die. This effect is large for populations with stars that have about the same age. The effect also increases with age of the whole population, since the differences in both the lifetimes and luminosities between the main sequence and the later stages increase with decreasing mass. For this work, we use the \\emph{Monte Carlo} method to bypass this problem while still avoiding the interpolation of tracks with its accompaining danger of creating artificial states. This is described in detail in section \\ref{sec_nummethod}. The star formation history of any stellar system can be described by a superposition of SSP models of different ages and metallicities. An example of this is given by Cellone \\& Forte et al (\\cite{cellone}) in their study of Low Surface Brightness galaxies or Contardo et al. (\\cite{contardo}) who investigate the formation and evolution of galaxies in a cosmological scenario. ", "conclusions": "In this work we present Monte-Carlo evolutionary synthesis models for SSPs which cover a wide range in metallicity, from $Z=0.0001$ to $Z=0.05$, using most recent and complete sets of input physics: stellar evolutionary tracks for stellar masses from $0.08 M_\\odot$ to $120 M_\\odot$, including post - helium flash evolution and mass loss, model atmosphere libraries also covering late stellar types and giving colours from $U$ to $K$ in agreement with observations and empirical calibrations for a series of absorption indices. We obtain theoretical calibrations of colours and indices in terms of metallicity which for model ages of 10-15 Gys agree closely with observations of GCs. The theoretical calibrations extend beyond the range of observed GCs, i.e. to a metallicity up to $[M/H] \\leq 0.4$. Moreover, our models provide these theoretical calibrations for all ages from cluster formation to 15 Gyrs and thus can also be applied in the interpretation of young star cluster systems observed in many interacting and starburst galaxies. The complete model files are available via WWW on {\\tt http://www.uni-sw.gwdg.de/\\~\\ okurth/ssp.html}. There are also models with different parameters for the IMF and for mass loss available." }, "9803/astro-ph9803100_arXiv.txt": { "abstract": "I present several simple figures to illustrate cosmology and structure formation in a nutshell. Then I discuss the following argument: if we assume that $\\Omega_{\\Lambda} = 0$ then the CMB results favor high $\\Omega_{m}$ while the supernova results favor low $\\Omega_{m}$. This large inconsistency is strong evidence for the incorrectness of the $\\Omega_{\\Lambda}=0$ assumption. Finally I discuss recent CMB results on the slope and normalization of the primordial power spectrum. ", "introduction": "The Big Bang model became the standard cosmological model soon after the discovery of the cosmic microwave background (CMB). The Big Bang model has a hot, dense early epoch (see Figure 1) when nucleosynthesis occurred. It also has an opaque surface that can naturally produce the Planckian spectrum of the CMB. The Steady State universe was not hotter in the past, has no epoch of Steady State Nucleosynthesis and has no opaque surface to produce the CMB. Gravitational collapse is the leading model of structure formation (Figure 2). Slight over-densities are gravitationally unstable and collapse under their own self-gravity. In an alternative family of models, structure forms from topological defects. In gravitational collapse models CMB anisotropies larger than $\\sim 1$ degree are acausal and rely on inflation to explain their existence. In defect models these large anisotropies are close by, causal and sub-horizon sized. \\thefirstfig \\clearpage \\thesecondfig \\setcounter{figure}{2} \\begin{figure}[hbt] \\centerline{\\psfig{file=wearehere.eps,height=70mm,width=110mm,angle=-90}} \\vspace{0cm} \\caption{Galaxies are CMB anisotropies are Quantum fluctuations. According to the inflationary scenario, quantum fluctuations of a scalar field are the origin of all structures. These quantum fluctuations are not caused by any preceeding event in the same sense as radioactive decay or quantum tunneling are not caused. They are non-deterministic prime movers. Inflation of the universe by a factor of more than $10^{26}$ transforms these quantum fluctuations into super-horizon classical density fluctuations. On their way to becoming galaxies we can monitor their progress by looking at CMB maps.} \\end{figure} \\clearpage One of the most important questions in cosmology is: What is the origin of all the galaxies, clusters, great walls, filaments and voids we see around us? The inflationary scenario provides the most popular explanation for the origin of these structures: they used to be quantum fluctuations. Figure 3 illustrates the metamorphosis of quantum fluctuations to CMB anisotropies to galaxies. Primordial quantum fluctuations of a scalar field get amplified and evolve to become classical seed perturbations and eventually large scale structure. This process can be monitored by CMB observations since matter fluctuations produce temperature fluctuations in the CMB: $\\frac{\\delta \\rho}{\\rho} \\propto \\frac{\\Delta T}{T}$. How does a particular fluctuation know whether it will become a spiral or an elliptical galaxy? Does the density and irregularity of its environment determine its morphology by controlling its angular momentum and the amount of merging? With a full understanding of galaxy formation we may be able to look at CMB cold spots and their neighborhoods and predict where they will end up in the Hubble tuning fork diagram of galaxy types. The distribution of morphological types at high redshift discussed by Driver in these proceedings would then be a derivable function of the characteristics of the CMB anisotropies. \\thefourthfig \\subsection{There is no scale beyond which the universe is homogeneous} It has been claimed that some recent, deep, galaxy redshift surveys have reached the scale at which the Universe becomes homogeneous. Strictly speaking however there is no scale beyond which the universe is homogeneous. The amplitude of the density contrast ($\\delta \\rho / \\rho \\propto k^{3}P(k)$ decreases for larger scales but is never zero. A more meaningful question is: Where is the turnover in the power spectrum? This turnover is due to a suppression of growth of a given k mode by $k^{4}$ relative to modes which enter the horizon during matter domination (assuming $\\oo = 1$). Thus, the horizon scale at matter-radiation equality is an important diagnostic of this fundamental scale. See Figure 4. Lineweaver \\& Barbosa (1998) have used current CMB anisotropy measurements to determine the position of the adiabatic peak in the CMB spectrum under the assumption of open or critical density CDM dominated universes: $\\ell_{peak} = 260^{+30}_{-20}$. Figure 5 illustrates how harmonic sound bumps appear in the CMB power spectrum driven by the wells and valleys of the CDM potentials. The epoch when matter and radiation densities are equal has a redshift of $z_{eq}$ while decoupling occurs at $z_{dec}$. The number of oscillations between $z_{eq}$ and $z_{dec}$ and thus the phase of the oscillations at $z_{dec}$ is determined by i) the physical size of the potential well, ii) the speed of sound and iii) the time interval between $z_{eq}$ and $z_{dec}$. \\vspace{0.7cm} \\thefifthfig \\clearpage ", "conclusions": "" }, "9803/astro-ph9803336_arXiv.txt": { "abstract": "This paper presents the preliminary results of the ESO Imaging Survey (EIS), a public survey being carried out by ESO and member states, in preparation for the VLT first-light. The survey goals, organization, strategy and observations are discussed and an overview is given of the survey pipeline developed to handle EIS data and produce object catalogs. A report is presented on moderately deep I-band observations obtained in the first of four patches surveyed, covering a region of 3.2 square degrees centered at $\\alpha \\sim 22^h 40^m$ and $\\delta =-40^\\circ$. The products available to the community, including pixel maps (with astrometric and photometric calibrations) and the corresponding object catalogs, are also described. In order to evaluate the quality of the data, preliminary estimates are presented for the star and galaxy number counts, and for the angular two-point correlation function obtained from the available data. The present work is meant as a preview of the final release of the EIS data that will become available later this year. ", "introduction": "With the advent of very large telescopes, such as the VLT, a largely unexplored domain of the universe becomes accessible to observations which may dramatically enhance on our understanding of different physical phenomena, in particular the origin and evolution of galaxies and large scale structures. In the next few years a wide array of 8-m telescopes will become available world-wide. Among these, the European VLT project is particularly striking because of its four 8-m telescopes and an impressive array of complementary instrumentation. Viewed as a unit, the VLT provides great flexibility by combining complementarity for certain programs with multiplexing capabilities for others. First-light for the VLT is scheduled for May 1998, with regular science operation starting in April 1999. In order to take full-advantage of the VLT from the start of its operation, ESO and its Observing Programmes Committee (OPC) decided to coordinate an imaging survey to provide candidate targets well-suited to the first set of VLT instruments. The ESO Imaging Survey (EIS) has been conceived as a collaborative effort between ESO and astronomers in its member states. Following the recommendation of the OPC, the survey has been overseen by a Working Group (WG). The EIS WG is composed of leading experts in different fields and has the responsibility of defining the survey science goals and strategy, and monitor its progress. In order to carry out the survey a dedicated team was assembled, starting March 1997. To stimulate cooperation between ESO and the astronomical community of the member states, EIS has sponsored the participation of experts as well as students and post-docs from the community in the development of software, observations and data reduction. As described by Renzini \\& da Costa (1997) (see also ``http://www.eso.org/eis''), EIS consists of two parts: EIS-wide to search for rare objects (\\eg distant clusters and quasars) and EIS-deep to define samples of high-redshift galaxies. These science goals were chosen to match as well as possible the capabilities of the first VLT instruments, FORS, ISAAC and UVES. EIS is also an essential first step in the long-term effort, currently underway at ESO, to provide adequate imaging capabilities in support of VLT science (Renzini 1998). The investment made in EIS will be carried over to a Pilot Survey utilizing the ESO/MPIA 2.2m telescope at La Silla, with its new wide-field camera. This Pilot Survey, which will follow the model of EIS, has been recommended by the EIS WG and is being submitted to the OPC. The goal of this paper is to describe the characteristics of I-band observations carried out in the fall of 1997 over a region of 3.2 square degrees (EIS patch A, da Costa \\etal 1998a) and of the corresponding data products, in the form of calibrated images and single frame catalogs. These products have been made publicly available through the ESO Science Archive, as a first step towards the full distribution of the EIS data. The purpose of the present release is also to provide potential users with a preview of the data, which may help them in the preparation of VLT proposals, and to encourage the community to provide constructive comments for the final release. It is important to emphasize that due to time limitations the results presented here should be viewed as preliminary and improvements are expected to be made before the final release of the EIS data later this year. In section 2, a brief description is presented of the criteria adopted in the field selection, the strategy of observations and the characteristics of the data in patch A already completed. It also describes the filters used, the definition of the EIS magnitude system and its relation to other systems, and the data used for the photometric calibration of the survey. In section 3, a brief description of the data reduction pipeline is presented, followed in section 4 by a description of the data products made publicly available in this preliminary data release. In section 5, the algorithms used to detect and classify objects, and the information available in the catalogs being distributed are described. Preliminary results from a scientific evaluation of the data is presented in section 6. In section 7, future plans are presented, followed in section 8 by a brief summary. ", "conclusions": "The ESO Imaging Survey is being carried out to help the selection of targets for the first year of operation of VLT. This paper describes the motivation, field and filter selection, and data reduction pipeline. Data for the first completed patch, in the form of astrometric and photometric calibrated pixels maps, single-frame catalogs, on-line coadded section images and further information on the project are available on the World Wide Web at ``http://\\-www.eso.org/eis''. Preliminary evaluation of the data shows that the overall quality of is good and the completeness limit of the extracted catalogs is sufficiently deep to meet the science requirements of EIS. Furthermore, the results for the other patches should improve as the observing conditions were considerably better than those in the period patch A was observed. The final and complete release of the data products of EIS is scheduled as follows: 1) EIS-wide, except the U-band : July 31, 1998, before the first call for proposals for the VLT; 2) EIS-deep and EIS-wide U-band on December 31, 1998." }, "9803/astro-ph9803046_arXiv.txt": { "abstract": " ", "introduction": "Thermal radiation from surfaces of several radio pulsars has been detected \\linebreak recently in soft X-rays by {\\it ROSAT\\/} and {\\it ASCA\\/} (see Becker and Tr\\\"umper,~1998). A number of the point-like X-ray sources has also been discovered and identified as radio-silent isolated cooling neutron stars (NSs) (see Caraveo et al.,~1996; Walter et al.,~1996; Vasisht et al.,~1997). In several cases, the observations seem to be better explained if the NSs possess an envelope of matter of low atomic weight, presumably hydrogen (Page,~1997), and if NS atmosphere models are applied to fit the data (Pavlov and Zavlin,~1998). There are indications that this may be the case for NSs with different magnetic fields $B$: from weak, $B\\ll10^{10}$~G (Rajagopal and Romani,~1996; Zavlin and Pavlov,~1998; Zavlin et al.,~1998), to strong, $B\\sim10^{11-13}$~G (Page et al.,~1996; Zavlin et al.,~1998), and superstrong, $B>10^{14}$~G (Heyl and Hernquist,~1997). To justfy this, one needs in further observations and advanced atmosphere models of NSs with high $B$. Thermal motion of atoms accross the magnetic field induces an electric field in the frame of the atom. This affects atomic structure, atmospheric thermodynamics and opacities (Ventura et al.,~1992; Pavlov and M\\'esz\\'aros,~1993), being a critical point in the construction of models and data interpretation. Here we study this problem for hydrogen plasma at temperatures $T\\sim 10^{5.5}-10^{6}$~K, densities $\\rho\\sim10^{-3}-10^1{\\rm~g~cm}^{-3}$, and magnetic fields $B\\sim 10^{12}-10^{13}$~G, typical for atmospheres of middle-aged cooling NSs. We construct an analytic model of the plasma free energy and derive a generalized Saha equation which is used to obtain the opacities. ", "conclusions": "" }, "9803/astro-ph9803270_arXiv.txt": { "abstract": "Warped \\ion{H}{1} gas layers in the outer regions of spiral galaxies usually display a noticeably twisted structure. This structure almost certainly arises primarily as a result of differential precession in the \\ion{H}{1} disk as it settles toward a preferred orientation in an underlying dark halo potential well that is not spherically symmetric. In an attempt to better understand the structure and evolution of these twisted, warped disk structures, we have adopted the ``twist--equation'' formalism originally developed by Petterson (1977) to study accretion onto compact objects. Utilizing more recent treatments of this formalism, we have generalized the twist--equation to allow for the treatment of non--Keplerian disks and from it have derived a steady--state structure of twisted disks that develops from free precession in a nonspherical, logarithmic halo potential. We have used this steady--state solution to produce \\ion{H}{1} maps of five galaxies (M83, NGC 300, NGC 2841, NGC 5033, NGC5055), which match the general features of the observed maps of these galaxies quite well. In addition, the model provides an avenue through which the kinematical viscosity of the \\ion{H}{1} disk and the quadrupole distortion of the dark halo in each galaxy can be quantified. This generalized equation can also be used to examine the time-evolutionary behavior of warped galaxy disks. ", "introduction": "In simplest terms, spiral galaxy disks can be described as geometrically thin, flat, and circular. We understand that spiral disks are geometrically thin because the gas of which they are composed is cold (the sound speed of the gas is much smaller than its circular orbital velocity); and they are both circular and flat because, being dissipative, the gas is fairly efficient at both minimizing out-of-the-plane motions and radial excursions that would lead to departures from circular orbits. Describing spiral disks as {\\it perfectly} circular and flat is clearly an oversimplification, however. In addition to the nonaxisymmetric structures that are obvious in optical photographs of many spiral disks, 21-cm maps of the projected velocity fields of spiral disks often reveal isovelocity contours that are significantly twisted (\\cite{RLW}; \\cite{RN78}; \\cite{RCC79}; \\cite{N80a}, 1980b; \\cite{B81}; \\cite{S85}). Kinematical tilted-ring models have been constructed in an effort to explain the presence of such twists in the velocity maps of \\ion{H}{1} disks. The models indicate that the outer regions of many normal spiral disks are significantly warped out of the principal plane that is defined by the optically visible, central portion of each galaxy. The line of nodes that defines the intersection of adjacent rings of gas in these kinematical models also usually must be twisted significantly as a function of radius in order to explain the observed contour maps (for recent reviews, see \\cite{Briggs90}; \\cite{Bosma91}; \\cite{CTSC93}, hereafter CTSC). It is not particularly surprising that many galaxies are observed to possess extended, rotationally flattened disks because such structures appear to be fairly ubiquitous in gravitationally bound, astrophysical systems. (There is strong evidence, for example, that rotationally flattened disks either exist now or have existed in the past around our sun, individual planets within our solar system, numerous protostars, the primary star of many mass--exchanging binary star systems, and active galactic nuclei.) What is peculiar about the \\ion{H}{1} disks of many galaxies is that the disks are significantly warped. It is not clear why natural dissipative processes similar to those which work effectively to minimize out-of-the-plane motions in stellar or protostellar ``accretion'' disks are unable to suppress warps in the gaseous disks of galaxies. As Binney (1992) has reviewed, there still is no generally accepted dynamical model of spiral galaxies that satisfactorily explains either the origin or the current structure of warped \\ion{H}{1} disks. Almost thirty years ago, Hunter \\& Toomre (1969) examined whether normal, infinitesimal bending oscillations might be exhibited by thin, rotating disks of self-gravitating material, permitting them to sustain coherent warps for more than a Hubble time. They concluded that ``for any disk whose density tapers sufficiently gradually to zero near its edge,'' the frequency spectrum of such oscillations is at least partly continuous and, as a result, coherent warps cannot be sustained. Over the subsequent decade, a number of other ideas surfaced to explain the persistence of warps in galaxies, each one taking advantage of the demonstrated existence of dark matter halos around galaxies. As Toomre (1983) has reviewed, however, models proposing to use the halo as an active agent to excite warps in otherwise flat disks -- for example, via the Mathieu instability (Binney 1981) or via a flapping instability (Bertin \\& Mark 1980) -- each present significant difficulties. Toomre (1983) and Dekel \\& Shlosman (1983) proposed, instead, that untwisted steady-state warps may be sustained as a result of steady forcing by a nonspherical, tilted halo. Building on the early work of Hunter \\& Toomre (1969) and the idea that forcing by a nonspherical dark halo can influence the dynamics of the visible disks of galaxies, Sparke (1986) and Sparke \\& Casertano (1988) have shown that a discrete warping mode can persist if the disk is sufficiently self-gravitating and if it is embedded in a nonspherical halo whose equatorial plane is tilted with respect to the centralmost regions of the disk. Adopting this model of galaxy warps, the following evolutionary picture emerges. During the galaxy formation process, gas which falls into a spheroidal dark matter halo generally will find that its angular momentum vector is tipped at some nonzero angle, $i$, away from the symmetry axis of the halo. If the gas is cold, it will settle into a rotationally flattened disk that is tilted at the same angle $i$ with respect to the equatorial plane of the halo. In the centralmost regions of the galaxy, where the self-gravity of the gas (or, ultimately, the combined gas/star system) dominates over the gravitational influence of the halo, the gas will be content to remain in orbits that preserve this original tilt. In the outermost regions of the galaxy where the gravitational field of the halo dominates, however, the gas should settle into the halo's equatorial plane. As Toomre (1983) and Dekel \\& Shlosman (1983) both sketched in their original concept papers, there also should be an intermediate region where the gas will be influenced significantly by the nonspherical gravitational fields of both the halo and the central gas (or gas/star) disk. Through their modeling efforts, Sparke (1986) and Sparke \\& Casertano (1988) confirmed this earlier suspicion that in the intermediate region, the gas can reside in a steady-state, ``warped disk'' structure that provides a smooth radial transition between the separate ``flat disk'' orientations of the inner and outer regions of the galaxy. Concentrating on the dynamics of the centralmost and intermediate regions -- that is, by building models in which there was effectively no gas in the outermost regions -- Sparke \\& Casertano (1988) showed that, in steady-state, the warped disk exhibits a straight line of nodes which precesses slowly and coherently in a direction retrograde to the orbital motion of the gas. In a time-dependent simulation, furthermore, Hofner \\& Sparke (1994) showed that settling to this steady-state warped disk structure occurs from the inside, out, and is driven not by dissipative processes that are similar to those which are thought to drive settling in most stellar or protostellar accretion disks but, rather, because ``bending waves carry energy associated with transient disturbances out toward the disk edge.'' They also showed that, during an evolution as the bending waves propagate outward through the intermediate region of the disk, a twisted structure can develop and persist until the gas has had sufficient time to settle into the steady-state (constant line of nodes) configuration. Sparke \\& Casertano (1988) and Hofner \\& Sparke (1994) have demonstrated that, with an appropriate choice of parameters, this model of disk warping matches well the observed properties of several galaxies with warped disks. (See also Kuijken 1991 and Dubinski \\& Kuijken 1995.) As Hofner \\& Sparke (1994) have pointed out, in galaxies with extended \\ion{H}{1} disks ``the outermost gas cannot be expected to form part of a coherent warping mode.'' They did not include normal dissipative forces in their simulations and therefore were unable to comment on how such forces might influence the settling process. In this paper, we examine the disk-settling process from the other extreme, ignoring the self-gravity of the gas but introducing an effective kinematical viscosity into the dynamical equations in order to simulate the effects of dissipative forces. Hence, our effort is complementary to the work of Hofner \\& Sparke (1994) and is most relevant to galaxies with extended \\ion{H}{1} disks -- although there is one galaxy used for model comparisons (NGC 2841) that is shared by both works. We adopt the view that warps in extended \\ion{H}{1} disks which exhibit substantial twists are transient features. Independent of precisely what physical process was responsible for initially placing the gas in an orbit that is inclined to the halo's equatorial plane ({\\it e.g.}, gas infall at the time of formation or a recent tidal encounter with another galaxy), the twisted structure can be understood as the result of differential precession in the gaseous disk as it dissipatively settles toward that ``preferred plane.'' In the past, there has been considerable concern (first enunciated by Kahn \\& Woltjer [1959], but reiterated in the reviews by both Toomre [1983] and Binney [1991]) that differential precession will destroy any warped disk structure on a time scale that is short compared to a Hubble time and, therefore, that the mechanism we are examining cannot reasonably be used to explain the persistence of such structures. In the outermost regions of \\ion{H}{1} disks, however, precession times are relatively long and, as was first pointed out by Tubbs \\& Sanders (1979), a warped gas layer can persist for a Hubble time if the dark halo in which the disk is embedded deviates only slightly from spherical symmetry. By modeling carefully the process of disk settling that is driven by normal dissipative forces and comparing the models to the observed kinematical properties of galaxies with extended, warped \\ion{H}{1} layers, we hope to be able to more carefully examine the viability of such models. Steiman-Cameron \\& Durisen (1988, hereafter SCD88) have developed this idea rather extensively. They have adopted a numerical, cloud-fluid model to simulate the time-dependent evolution of a galaxy disk that intially is tilted out of the equatorial plane of an underlying, spheroidal dark halo. The disk is assumed to be composed of a set of annular mass elements, or ``clouds,'' which act like atoms in a viscous fluid. The SCD88 model has offered some valuable physical insight into the time-dependent settling process that is driven by normal dissipative forces and their dynamically generated model of a twisted galaxy disk has been surprisingly successful at matching the peculiar optical image of one particular galaxy, NGC 4753 (\\cite{SCKD92}). Our model is analogous to the one developed by SCD88 but it derives from an analytical prescription of the viscous settling process. More specifically, we employ the ``twisted-disk'' equation formalism first developed by Bardeen \\& Petterson (1975) and Petterson (1977, 1978) to describe the time-dependent settling of a thin, viscous disk in a nonspherical dark halo potential. This is a rather natural formalism to adopt because, as numerous kinematical ``tilted-ring'' models have demonstrated, warped \\ion{H}{1} galaxy disks display a structure that resembles, at least qualitatively, the twisted geometry that had once been thought to be important in accretion disks which surround certain compact stellar objects (Bardeen \\& Petterson 1975; see a recent rejuvenation of this idea put forward by Maloney \\& Begelman 1997). In adapting the model to galaxy disks, we have replaced the approximate Keplerian gravitational potential used in earlier accretion disk work with a logarithmic potential appropriate to galaxy halos. (Pringle [1992] also recently described how the twisted--disk formalism may be adapted to galaxies.) In the limit of stress-free precession, our model reproduces the analytical prescription of disk settling first presented by SCD88, but our model is not constrained to this limit. A more general solution to the governing equations predicts an exponential settling rate that depends on time to the first power, rather than on time to the third power as has been derived in the limit of stress-free precession. Furthermore, an analytical, steady-state solution to the governing equations produces a twisted-disk structure that is very similar to previously constructed, kinematical models. We demonstrate that projected surface density maps and radial velocity maps derived from our analytical model match published \\ion{H}{1} maps of five well-studied warped disk galaxies (M83, NGC 300, NGC 2841, NGC 5033, and NGC 5055) very well. ", "conclusions": "Utilizing the formalism originally introduced by Petterson to describe warped and twisted accretion disk structures in Keplerian potentials, we have derived a single, complex ODE to describe time dependent settling of an \\ion{H}{1} disk in the logarithmic potential that appears to be typical of normal spiral galaxies. Over the interval $1 \\lesssim x \\lesssim 20$, the analytical function $w_{ss}\\bigl(x\\bigr)$ -- derived from an analysis of the steady-state limit of the general twisted-disk equation -- appears to describe quite accurately the general warped and twisted structure that is exhibited by a number of galaxy disks. It should be noted again that our analysis is based on the assumption that the warping angle $\\beta$ is $\\ll 1$. For galaxies with larger warps (including some that we have modeled) this is a simplifying assumption and should be disregarded in more proper treatments (see Pringle [1992]). From our model fits we conclude that, quite generally, the effective kinematical viscosity in these neutral hydrogen disks is $\\nu\\sim0.6$ km s$^{-1}$ kpc. That is, the effective Reynolds number in these systems is \\begin{displaymath} R_e\\sim\\frac{r_{max} v_{\\psi}}{\\nu} \\sim \\frac{x_{max}^2 v_\\psi}{r_{max}} \\sim6000\\;. \\end{displaymath} According to equation (\\ref{radvsub}), this also implies that the ratio of the radial inflow velocity of the gas to its orbital velocity is \\begin{displaymath} \\frac{v_r}{v_{\\psi}}\\approx\\frac{1}{R_e}\\sim\\times{10}^{-4}\\;. \\end{displaymath} This model also provides a mechanism by which the parameter $\\eta$ -- the quadrupole moment of the underlying dark halo potential well -- can be measured in spiral galaxies. Our fits to five normal spirals with well-studied warped \\ion{H}{1} disks specifically indicate that $\\eta\\sim{10}^{-3}$. (This value can be increased to $\\eta\\sim{10}^{-2}$, with a corresponding factor of ten increase in viscosity, only if the age of these disks is assumed to be 0.1$\\,{H_0}^{-1}$ -- which seems unlikely -- or $\\sigma$ in expression 38c is found to be $\\sim 0.1$.) Hence, we conclude that the dark halos in which these warped disks sit are, to quite high accuracy, spherically symmetric. This particular conclusion should not come as a surprise because some time ago Tubbs and Sanders (1979) pointed out that if warped disks are identified as transient structures, the warp can only be sustained for a Hubble time if the underlying halo potential is very nearly spherical. Although we have examined in detail here only the steady-state solution, the derived time-dependent twisted-disk equation provides a tool that can be utilized to model the time-evolution of warped \\ion{H}{1} disks without resorting to elaborate numerical techniques. The twisted-disk formalism in general -- and the analytical function $w_{ss}\\bigl(x\\bigr)$ in particular -- provides an avenue through which models of warped \\ion{H}{1} disks can advance from purely kinematical fits (e.g. tilted-ring models) to dynamical models based on a reasonable physical model. In the future, we also expect to use the time-dependent form of the twisted-disk equations as a point of comparison for fully three-dimensional gas dynamic simulations of \\ion{H}{1} disks settling into distorted halo potentials." }, "9803/astro-ph9803028_arXiv.txt": { "abstract": "We present a total of 48~minutes of observations of the nearby, bright millisecond pulsar \\psr~taken at the Parkes radio observatory in Australia. The data were obtained at a central radio frequency of 1380~MHz using a high-speed tape recorder that permitted coherent Nyquist sampling of 50 MHz of bandwidth in each of two polarizations. Using the high time resolution available from this voltage recording technique, we have studied a variety of single-pulse properties, most for the first time in a millisecond pulsar. We show that individual pulses are broadband, have pulse widths ranging from $\\sim$10~$\\mu$s ($\\sim 0.6^{\\circ}$ in pulse longitude) to $\\sim$300~$\\mu$s ($\\sim 20^{\\circ}$) with a mean pulse width of $\\sim$~65$\\mu$s ($\\sim 4^{\\circ}$), exhibit a wide variety of morphologies, and can be highly linearly polarized. Single pulse peaks can be as high as 205~Jy (over $\\sim$40 times the average pulse peak), and have a probability distribution similar to those of slow-rotating pulsars. We observed no single pulse energy exceeding $\\sim$4.4 times the average pulse energy, ruling out ``giant pulses'' as have been seen for the Crab and \\ntts\\ pulsars. \\psr\\ does not exhibit classical microstructure or show any signs of a preferred time scale that could be associated with primary emitters; single pulse modulation has been observed to be consistent with amplitude-modulated noise down to time scales of 80 ns. We observe a significant inverse correlation between pulse peak and width. Thus, the average pulse profile produced by selecting for large pulse peaks is narrower than the standard average profile. We find no evidence for ``diffractive'' quantization effects in the individual pulse arrival times or amplitudes as have been reported for this pulsar at lower radio frequency using coarser time resolution (\\cite{amdv97}). Overall, we find that the single pulse properties of \\psr\\ are similar to those of the common slow-rotating pulsars, even though this pulsar's magnetosphere and surface magnetic field are several orders of magnitude smaller than those of the general population. The pulsar radio emission mechanism must therefore be insensitive to these fundamental neutron star properties. ", "introduction": "\\label{sec:intro} Single pulse studies of millisecond pulsars are considerably more difficult to perform than are those of slow pulsars. Faster data rates are needed to study millisecond pulsars with comparable pulse phase resolution, and finer radio frequency resolution is required to minimize the effect of interstellar dispersion. Also, interstellar scattering time scales comparable to the pulse duration render studying individual pulse morphologies impossible, while steep millisecond pulsar spectra preclude observations of sufficient sensitivity at higher radio frequencies where scattering is less important. Yet single pulse studies of millisecond pulsars are highly desirable for two reasons. First, the origin of the radio emission that has made isolated neutron stars famous is, even 30 years after their discovery, still a mystery. The high brightness temperatures ($\\sim 10^{30}$K) associated with the radio emission point to coherent processes which are poorly understood even under less exotic conditions (\\cite{mel96}). Previous observations of slow pulsars have not constrained the emission mechanism sufficiently; the study of radio emission properties of millisecond pulsars may provide important new clues. Millisecond pulsars, because of their fast spin periods, have much smaller light-cylinder radii, and hence magnetospheres, than slow pulsars. They also have lower surface magnetic field strengths than the general pulsar population (most likely resulting from their having been ``recycled'' by a binary companion through the accretion of mass and angular momentum). Were the radio emission mechanism at all dependent on such properties, millisecond pulsars should have different radio properties than the slower-spinning general population. The second reason single pulse studies of millisecond pulsars are important is that millisecond pulsar timing is well-known to be an unparalleled source of precision astrometric and astrophysical information. Among factors possibly limiting timing precision is the stability of the average profile, which depends on the properties of single pulses. Only recently has a systematic study of single pulses from millisecond pulsars become possible, largely due to improving computational and data recording technologies. To date, the only such investigation has been for the 1.5~ms pulsar \\ntts\\ (\\cite{wcs84,sb95,bac95,cstt96}). Interestingly, \\ntts\\ exhibits giant radio pulses like those seen elsewhere only in the Crab pulsar. With the single pulse properties of only one millisecond pulsar having been studied in any detail, the question of whether all millisecond pulsars show similar properties naturally arises. Unfortunately, \\ntts\\ suffers interstellar scattering at time scales comparable to the duration of a single pulse at the radio frequencies at which it has been observed, rendering detailed study of individual pulse morphologies difficult. Here we report on high-time-resolution single pulse studies of a second millisecond pulsar, \\psr. This pulsar's large flux density and low dispersion measure (DM), and the corresponding scarcity of line-of-sight scattering material, render it an obvious target for single pulse work. Some single pulse investigations of \\psr\\ have been reported (\\cite{jlh+93,amdv97}) but none have had sufficient time resolution to resolve most individual pulses. Using a fast recording device and powerful supercomputers, we have been able to resolve all pulses in our data, the narrowest being $\\sim$10~$\\mu$s. ", "conclusions": "\\label{sec:concl} We have presented the first detailed single pulse study of a millisecond pulsar in which sufficient time resolution was available to resolve single pulses as they would be seen in the pulsar vicinity. The similarity of the single pulse properties to those of normal slow pulsars is remarkable given the dramatically reduced magnetospheric volume and magnetic field strength of \\psr. Indeed, without being told the absolute sample rate, it is unlikely that one could distinguish between this being a millisecond or slow pulsar. To summarize, our observations of \\psr\\ have: resolved individual pulses, and shown that they have a wide variety of morphologies, with multiple sub-pulse components not uncommon; shown that individual pulses are in general broadband; shown that individual pulses can have high linear polarization; provided no evidence for giant pulses as observed in the Crab pulsar and PSR B1937+21, nor for pulse nulling or drifting subpulse phenomena; found structure in the intensity fluctuation spectrum; revealed a correlation of pulse peak with pulse width so that the average profile formed from only the highest amplitude pulses is much narrower than the conventional average profile; not shown any evidence for micro-structure or preferred time scales $\\geq$80~ns; shown that the emission is consistent with an amplitude modulated noise model; provided no evidence to support the claim made by Ables \\etal\\ (1997) of the detection of coherent radiation patterns. Because there exists no self-consistent radio emission mechanism context (see e.g. \\cite{mel96}) in which to discuss these results, it is difficult to say how such models are constrained. Indeed previous slow pulsar single pulse studies suffered from the same difficulty. However, it is often the case that fundamental insights become apparent when known phenomenon are taken to extremes; this, and improved tape recording and computer technologies that permit single pulse studies of millisecond pulsars, provided our motivation in undertaking this analysis. Similar studies of other millisecond pulsars may eventually lead to an understanding of the radio emission mechanism." }, "9803/astro-ph9803264_arXiv.txt": { "abstract": "We present multi-frequency radio continuum VLBI observations of the gravitational lens system B0218+357 carried out using a global VLBI network and the VLBA. The source has been observed with resolutions from 0.2~mas to 5~mas and displays interesting structure. The spectral properties of various components show that the lensed object is a standard flat spectrum radio source which has many self-absorbed components. Based on the flux ratio of the lensed images as a function of frequency we propose a simple model for the background radio source. ", "introduction": "The radio source B0218+357 has been identified as a gravitationally lensed system (Patnaik et al. 1993). The source, which has a core-halo structure at low frequency (1.4~GHz) and low resolution (5~arcsec), consists of two compact flat-spectrum components, A and B, separated by 335 milliarcsec. The weaker of the two, B, is surrounded by a faint Einstein ring of similar diameter. The spectral and polarisation characteristics of the components show them to be the lensed images of a single flat spectrum `core'. The lensing galaxy, suggested to be a spiral galaxy, has been observed by the HST (Jackson et al. 1997). It has a redshift of 0.6847 (Browne et al. 1993). Atomic hydrogen and many molecular species have been detected in absorption against the background source (Carilli et al. 1993, Wiklind \\& Combes 1995, Menten \\& Reid 1996). The redshift of the background object is suggested to be 0.96 (Lawrence 1996). In this contribution we explore the properties of the radio source from our multi-frequency VLBI observations. We give a summary of our observations and discuss the results in the context of a flat spectrum radio source. We note that the overall radio spectrum of B0218+357 is flat between 365~MHz and 43~GHz i.e. its flux density is constant within a factor of two between these frequencies. ", "conclusions": "We derive two basic results from these observations, namely the spectral decomposition of various components and the ratio of their flux densities as a function of frequency. Since we resolve the `core-jet' structure in A and B at frequencies higher than 8.4~GHz, we have 4 measurements of flux density for these components. The spectra are plotted in Fig. 2. The spectrum of the core (A1 and B1) is self-absorbed with a peak around 15~GHz. The jet (A2 and B2) has steep spectra above 8.4~GHz. However, the total flux density of A and B at 1.7 and 5~GHz suggest that both A2 and B2 must also be self-absorbed between these two frequencies. The Einstein ring emission has a steeper spectrum between 5 and 22~GHz and it must also be self-absorbed at lower frequencies since the total flux density provides a limit (Patnaik et al. 1993). In summary, the radio structure of B0218+357 consists of at least 3 distinct components, core, jet and the Einstein ring, which are all self-absorbed at different frequencies. The components are self-absorbed at progressively lower frequencies as one moves from the core along the jet. This behaviour of the radio source is consistent with its flat spectrum (Cotton et al. 1980). The core, being self-absorbed around 15~GHz, contributes very little to the emission at 1.7 and 5~GHz, where the emission is dominated by the jet as evident from the spectra. Perhaps it is not surprising therefore that the radio structures of both A and B at 1.7~GHz appear diffuse. However, this poses a difficulty in that a self-absorbed component, which must be compact, has a large observed size. In this case one must consider the intrinsic size of the source rather than the observed size which has been magnified by the lens. Another result from our observations is that the image flux ratios change with frequency. This is an apparent contradiction since gravitational lensing is achromatic. However, since the source is extended and spans areas of different image magnifications the observed flux ratio can indeed vary with frequency. The flux ratio of the core (i.e. A1/B1) is around 3.7 for frequencies higher than 8.4~GHz. This is similar to the ratio of A/B obtained from VLA observations at these frequencies (Patnaik et al. 1993). At lower frequencies the flux ratio A/B is around 2.6. Of course, we point out that the source is variable at high frequencies and thus the measured flux ratio can be different due to differential time delay. The above results lead to a rather simple model for B0218+357. The radio source consists of a number of different components which are self-absorbed at different frequencies thus conspiring to produce the observed flat spectrum. The core (imaged into A1 and B1) is located at the western-most edge, the jet (imaged into A2 and B2) lies between the core and the component giving rise to the Einstein ring. The core dominates the radio structure at frequencies higher than 8.4~GHz, and the jet and the Einstein ring dominate at lower frequencies. At lower frequencies one does not detect any compact feature in the source since the core contributes very little to the flux density." }, "9803/astro-ph9803322_arXiv.txt": { "abstract": "We have determined detailed radio spectra for 26 compact sources in the starburst nucleus of M82, between $\\lambda\\lambda$ 74 and 1.3 cm. Seventeen show low-frequency turnovers. One other has a thermal emission spectrum, and we identify it as an \\HII region. The low frequency turnovers are due to absorption by interstellar thermal gas in M82. New information on the AGN candidate, 44.01+595, shows it to have a non-thermal falling powerlaw spectrum at the highest frequencies, and that it is strongly absorbed below 2 GHz. We derive large magnetic fields in the supernova remnants, of order 1-2 $(1+k)^{2/7} \\phi^{-2/7}$ milliGauss, hence large pressures in the sources suggest that the brightest ones are either expanding or are strongly confined by a dense interstellar medium. From the largest source in our sample, we derive a supernova rate of 0.016 yr$^{-1}$. ", "introduction": "Supernovae (SNe) and supernova remnants (SNR's) are thought to be the main ``drivers'' of the starburst phenomenon (\\cite{kw75}). Detailed radio observations for supernovae and supernova remnants are few, due to the fact that only a small subset of the extragalactic SNe discovered each year produce detectable radio emission. Those events for which monitoring data is available (e.g. \\cite{weiler88}) have revealed considerable information regarding the evolution of the properties of the expanding shock wave and its emission processes. In intense starbursts like M82, where a significant population of massive stars has had sufficient time to evolve to the SN stage, we have the unique opportunity to study a collection of SNR's of similar age and origin, and their interaction with the surrounding environment. The observations are briefly outlined in Section 2. We discuss the technique used to fit spectral models in Section 3. Then in Section 4 we discuss the physical implications of our modelling. Throughout the paper, we assume that the distance to M82 is 3.63 Mpc (\\cite{freedman}, \\cite{tammann}), so that 1$\\farcs$0 corresponds to 17.6 pc. ", "conclusions": " 1. Of the 26 sources, one (42.21+590) is an \\HII region, 22 are most likely young radio supernovae or supernova remnants, and three further sources remain of uncertain nature. 2. High magnetic field strengths and pressures are derived for the sources we identify as SNR's. The brighter ones do not appear to be in pressure equilibrium with the surrounding medium, and are probably expanding, or are strongly confined by the high pressure ISM. The source of the high nuclear pressure and halo winds are most likely the SNR's. 3. Sources that show a low-frequency turnover, do so as a result of absorption by ionized gas. This ionized gas may be intimately associated with the source, or may be in clouds located along the line of sight to the sources. The gas is most likely in the form of clumpy, high density clouds. 4. The resolved ring-like source in our survey is the oldest. We derive a supernova rate in M82 of $\\sim$ 0.016 $\\left( \\frac{V}{5000 \\;\\;km/s} \\right)$yr$^{-1}$. This value is a lower limit." }, "9803/astro-ph9803114_arXiv.txt": { "abstract": "As the Cosmic Microwave Background (CMB) radiation is observed to higher and higher angular resolution the size of the resulting datasets becomes a serious constraint on their analysis. In particular current algorithms to determine the location of, and curvature at, the peak of the power spectrum likelihood function from a general $N_{p}$-pixel CMB sky map scale as $O(N_{p}^{3})$. Moreover the current best algorithm --- the quadratic estimator --- is a Newton-Raphson iterative scheme and so requires a `sufficiently good' starting point to guarantee convergence to the true maximum. Here we present an algorithm to calculate bounds on the likelihood function at any point in parameter space using Gaussian quadrature and show that, judiciously applied, it scales as only $O(N_{p}^{7/3})$. Although it provides no direct curvature information we show how this approach is well-suited both to estimating cosmological parameters directly and to providing a coarse map of the power spectrum likelihood function from which to select the starting point for more refined techniques. ", "introduction": "Planned observations of the Cosmic Microwave Background (CMB) will have sufficient angular resolution to probe the CMB power spectrum up to multipoles $l \\sim 1000$ or more (for a general review of forthcoming observations see \\cite{S}). If we are able to extract the multipole amplitudes $C_{l}$ from the data sufficiently accurately we will be able to obtain the values of the fundamental cosmological parameters to unprecedented accuracy. The CMB will then have lived up to its promise of being the most powerful discriminant between cosmological models \\cite{Kn,HSS,KKJS}. Extracting the power spectrum is conceptually simple --- the raw data is cleaned and converted into a time-ordered dataset. This is then converted to a sky temperature map, which in turn is analysed to find the location of, and curvature at, the maximum of the likelihood function of the power spectrum. In practice as the size of the dataset increases the problem rapidly becomes intractable by conventional methods. This is particularly true of the final step --- the likelihood analysis of any reasonably general sky temperature map. An observation of the CMB contains both signal and noise \\begin{equation} \\Delta_{i} = s_{i} + n_{i} \\end{equation} at each of $N_{p}$ pixels. For independent, zero-mean, signal and noise the covariance matrix of the data \\begin{equation} M \\equiv \\left< {\\bf \\Delta} \\, {\\bf \\Delta}^{T} \\right> = \\left< {\\bf s} \\, {\\bf s}^{T} \\right> + \\left< {\\bf n} \\, {\\bf n}^{T} \\right> \\equiv S + N \\end{equation} is symmetric, positive definite and dense. For any theoretical power spectrum $C_{l}$ we can construct the corresponding signal covariance matrix $S(C_{l})$; knowing the noise covariance matrix $N$ for the experiment we now know the observation covariance matrix for that power spectrum $M(C_{l})$. The probability of the observation given the assumed power spectrum is then \\begin{equation} \\label{eq.lf} P({\\bf \\Delta} \\, | \\, C_{l}) = \\frac{e^{-{\\small\\frac{1}{2}} \\, {\\bf \\Delta}^{T} M^{-1} {\\bf \\Delta}}} {(2 \\pi)^{N_{p}/2} \\left| M \\right|^{1/2}} \\end{equation} Assuming a uniform prior, so that \\begin{equation} P(C_{l} \\, | \\, {\\bf \\Delta}) \\propto P({\\bf \\Delta} \\, | \\, C_{l}) \\end{equation} we can restrict our attention to evaluating the right hand side of equation (\\ref{eq.lf}). Unfortunately unless the noise covariance matrix is unrealistically simple (eg. diagonal) both direct evaluation \\cite{G1,G2} and quadratic estimation \\cite{T,BJK} of the likelihood function scale as at best $O(N_{p}^{3})$ \\cite{B}, making them impractical for the forthcoming $10^4$ -- $10^6$ pixel datasets. ", "conclusions": "We have presented an algorithm to calculate probabilistic bounds on the power spectrum likelihood function from an $N_{p}$-pixel CMB map using Gaussian quadrature which scales as between $O(N_{p}^{7/3})$ and $O(N_{p}^{5/2})$ --- a very significant advance on existing algorithms for the exact calculation which scale as $O(N_{p}^{3})$. Since lowering the convergence constraint below the 1\\% level gains us only marginally tighter final bounds at the expense of increasing the scaling power, it is not recommended for the forthcoming $10^{4}$ -- $10^{6}$ pixel CMB maps. Our final algorithm of choice therefore gives better than 3\\% bounds on the logarithm of the likelihood function with $O(N_{p}^{7/3})$ operations with 99\\% confidence. Since this algorithm gives no information about the local curvature of the likelihood function it is not as well suited as quadratic estimator techniques for searching a large multi-dimensional parameter space for its likelihood maximum. However for direct estimation of a small set of cosmological parameters this technique is certainly viable and fast. Moreover, even when the parameters are taken to be the multipole moments (individually or in bins), quadratic estimation, being a Newton-Raphson iteration, requires a starting point `sufficiently close' to the maximum to guarantee convergence; the algorithm presented here is then well suited to provide a coarse overall mapping of the likelihood function from which to select a starting point for more refined techniques." }, "9803/astro-ph9803232_arXiv.txt": { "abstract": "We study the influence of a possible H dibaryon condensate on the equation of state and the overall properties of neutron stars whose population otherwise contains nucleons and hyperons. In particular, we are interested in the question of whether neutron stars and their masses can be used to say anything about the existence and properties of the H dibaryon. We find that the equation of state is softened by the appearance of a dibaryon condensate and can result in a mass plateau for neutron stars. If the limiting neutron star mass is about that of the Hulse-Taylor pulsar a condensate of H dibaryons of vacuum mass $\\sim 2.2$ GeV and a moderately attractive potential in the medium could not be ruled out. On the other hand, if the medium potential were even moderately repulsive, the H, would not likely exist in neutron stars. If neutron stars of mass $\\sim 1.6 M_\\odot$ were known to exist, attractive medium effects for the H could be ruled out. ", "introduction": "Since Jaffe proposed that there may exist a stable dihyperon (a quark composite with baryon number two) \\cite{Jaffe77}, an ongoing quest for this particle began \\cite{Carroll78}. Recent searches using kaon beams \\cite{Aoki90} or heavy ion beams \\cite{Belz96,Belz97,Stotz97} found no candidates or are still in progress \\cite{Craw98}. There exist some claims for evidence for the H dibaryon produced in proton-nucleus \\cite{Shahba93} and in heavy-ion collisions \\cite{Long95}. Nevertheless, these candidates might be misidentified $K^0_L$ as seen in \\cite{Belz97}. For a most recent overview on the search for the H dibaryon we refer to \\cite{HYP97}. There are numerous mass estimates for the H dibaryon and they are reviewed in \\cite{Dover89}. The existence or nonexistence of the H dibaryon is strongly connected with the observation of double $\\Lambda$ hypernuclei which has been discussed in \\cite{Dalitz89}. Three double $\\Lambda$ hypernuclei have been reported in literature: $^{~~6}_{\\Lambda\\Lambda}$He \\cite{Danysz63}, $^{~10}_{\\Lambda\\Lambda}$Be \\cite{Prowse66}, and $^{~13}_{\\Lambda\\Lambda}$B \\cite{Aoki91,Dover91}. The two $\\Lambda$'s can decay by strong interactions to the H dibaryon. As this has not been seen in the above events, the H must either be heavier than $m_H> 2 m_\\Lambda+B_{\\Lambda\\Lambda}\\approx 2.22$ GeV \\cite{Kerb84} or the events are misidentified as an H hypernucleus with a shallow attractive nuclear potential \\cite{Dover89}. A more stringent condition is the observation of the weak mesonic decay of the double $\\Lambda$ hypernuclei giving $m_H> m_\\Lambda + m_p + m_{\\pi^-} + B_\\Lambda \\approx 2190$ MeV \\cite{Jaffe91} where $B_\\Lambda$ depends on the mass of the decay fragment and is $B_\\Lambda=-3.1$ MeV for $^5_\\Lambda$He. In all cases, a deeply bound H dibaryon seems to be ruled out by these events. If the H dibaryon exists, it will have a certain impact also on the properties of dense matter. It is quite established nowadays, that neutron stars have a large hyperon fraction in the core and might be described as giant hypernuclei, though bound by gravity \\cite{Glen85}. Here again, the presence of hyperons might restrict certain properties of the H dibaryon. Recently, studies for neutron stars have been done for nuclear matter without hyperons but including H dibaryon condensation \\cite{Fae97a} and limits have been set for the coupling constants of the H dibaryon \\cite{Fae97b}. There might exist heavier partners of the H dibaryon, lumps of strange quark matter dubbed strangelets. There are several heavy-ion experiments dedicated to search for this novel form of matter \\cite{Arm97,Beavies95,Apple96}. In the MIT bag model, strangelets with $A\\leq 6$ are found to be unbound \\cite{Aerts78}. Nevertheless, light strangelet candidates in the range of $63$) narrow H$\\beta$ line, very strong FeII emission and shows X-ray variations of a factor 50 or more on timescales of a few days (Otani 1995, Brandt et al. 1995, Boller et al 1997) and a factor of 2 in less than 800 seconds (Boller et al. 1996); \\item RE J 1237+264 showed one very large (factor of 50) variability event (Brandt, Pounds \\& Fink, 1995) \\item PHL1092 varied by a factor of 4 in 2 days (Forster \\& Halpern 1996, Lawrence et al. 1997). PHL1092 is a high luminosity ($5\\times10^{46}$ erg s$^{-1}$) very steep 0.1-2 keV spectrum and narrow emission line quasar. \\item The relatively high luminosity quasar NAB0205+024 ($L_{0.5-10 keV}=8\\times10^{44}$ erg s$^{-1}$) shows variations of a factor of 2 in less than 20 ks in an ASCA observation (Fiore et al. 1998a). \\item Mark 478 (PG1440+356) shows factor of 7 variations in 1 day during a long EUVE monitoring (Marshall et al 1996) \\item The extremely soft NLSy1 WPVS007 ($\\alpha_{0.1-2keV}=7.3$) showed a huge variation (factor of 400) variation between the RASS observation and a follow-up PSPC pointed observation taken 2 years later (Groupe et al. 1995). \\end{enumerate} At this point we believe that a systematic study of a sample of normal quasars spaning the observed range of properties is needed. A systematic study of the variability properties of AGNs is not an easy task, since it requires: \\begin {enumerate} \\item the selection of well defined and representative (i.e. unbiased) sample; \\item the availability of numerous repeated observations; \\item a carefully designed observational strategy. In particular, it is very important that the sampling time is regular and that it is similar for all objects in the sample, so that the results on each object can be easily compared with each other. Otherwise the differences in the observed variability may be induced just by differences in sampling patterns. \\end{enumerate} Because of these stringent requirements systematic studies are still lacking (for example the Zamorani et al. 1984 study on quasars, and the Green et al. 1993 study on Seyfert galaxies do not fulfill any of the three criteria above). To explore in a systematic manner the possible relation between line width, soft X-ray spectrum and X-ray variability properties, we initiated a pilot program using the ROSAT HRI. In the following we present the results from a campaign of observations of six PG quasars which addresses the above points. ", "conclusions": "In our sample of six PG quasars we found clear evidence that X-ray steep quasars show larger amplitude variations than X-ray flat quasars on timescales from 2 days to 20 days. On longer timescales we do not find significant differences between steep and flat quasars, although the statistics are poorer. While the sample in this pilot study is small (and expanded monitoring of a larger sample of quasars is surely needed), the distinction is so clean cut that we feel justified in speculating on its origin. Large narrow band flux variability in sources with a steep spectrum could be induced by small changes in the spectral index without large variations of the spectrum normalization. If this is the case each source should show large spectral variability, with $\\alpha_X$ anti-correlated (correlated) with the observed flux if the pivot is at energies lower (higher) than the observed band. Spectral variability of this kind has not been observed in the PSPC observations of these and other quasars (Laor et al. 1997, Fiore et al. 1994) and in NLSy1 (Boller et al 1997, Fiore et al. 1998a, Brandt et al. 1995). Although our HRI observations cannot directly rule out this possibility, we therefore conclude that the observed temporal variability pattern is not likely the result of complex spectral variability. Laor et al (1997) suggest that a possible explanation for the remarkably strong $\\alpha_X$--$H_{\\beta}$ FWHM correlation is a dependence of $\\alpha_X$ on $L/L_{\\rm Edd}$. The line width is inversely proportional to $\\sqrt {L/L_{\\rm Edd}}$ if the broad line region is virialized and if its size is determined by the central source luminosity (see Laor et al. 1997, \\S 4.7). So narrow-line, steep (0.1-2 keV) spectrum AGNs emit close to the Eddington luminosity and have a relatively low mass black hole. A similar dependence of spectral shape on $L/L_{\\rm Edd}$ is seen in Galactic black hole candidates (BHC) as they change from the `soft-high' state to the `hard-low' state. A physical interpretation for this effect, as described by Pounds et al. (1995), is that the hard X-ray power-law is produced by Comptonization in a hot corona and that as the object becomes more luminous in the optical-UV, Compton cooling of the corona increases, the corona becomes colder, thus producing a steeper X-ray power-law. In BHC in `soft-high' states this power law component emerges above $\\sim 10$ keV, while the spectrum below this energy is dominated by softer emission often associated with optically thick emission from an accretion disk. In this model quasar disc emission is at a too low an energy to observe, since the disc temperature scales with the mass of the compact object as $M_{BH}^{-1/4}$, leaving the Comptonized power law to dominate the 2-10 keV spectrum. The 0.2-2 keV quasar emission could be due to Comptonization (see e.g. Czerny \\& Elvis 1987 and Fiore et al. 1995) by a second cooler gas component. The 0.2-2 keV and 2-10 keV spectral indices might well be correlated with each other, if the emitting regions are connected or if the emission mechanisms know about each other. We have undertaken a campaign of observations of the quasars in the Laor et al. (1997) sample with ASCA and BeppoSAX to clarify this point. Preliminary results (Fiore et al. 1998b) shows that this is indeed the case: steep $\\alpha_X$(PSPC) quasars tends to have a steeper hard (2-10 keV) X-ray power-law (although the spread of the 2-10 keV indices seems smaller than that of the PSPC indices (as also found by Brandt et al. 1997). In a sample of quasars with similar luminosities, those emitting closer to the Eddington luminosity will also be those with the smaller black hole and hence smaller X-ray emission region. Thus light travel time effects would smear intrinsic X-ray variability up to shorter time scales in high $L/L_{Edd}$ objects, compared with low $L/L_{Edd}$ objects. Based on this interpretation, and on figures \\ref{stru_med} and \\ref{stru_mean} it appears that the emission region of the steep soft X-ray quasars is a factor of $\\approx 10$ smaller than that of flat soft X-ray quasars (about $\\approx 10^{16}$ cm and $10^{17}$ cm respectively). Alternatively, the higher variability of steep spectrum objects may be a true intrinsic property, which could be induced by some increased instability in high $L/L_{Edd}$ objects. The range of luminosity in our sample is too small to tell if the variability amplitude appears to be better correlated with black hole mass, as expected for light travel time smearing, or with $L/L_{Edd}$, which would indicate an intrinsic mechanism. It is interesting to note that a completely different analysis, based on the interpretation of the optical to X-ray spectral energy distribution in terms of emission from accretion disks also suggest a small black hole mass and an high arretion rate in two NLSy1 (Siemiginowska et al. 1998). It is also interesting to note that Cyg X-1 in the ``high and soft'' state (Cui et al. 1997) shows a total root mean square variability higher than that measured during periods of transitions and in the ``low and hard'' state. This is due to strong 1/f noise, extending down to at least a few $10^{-3}$ Hz, when the source is the ``high and soft'' state (Cui et al. 1997). When the source is in the ``low and hard'' state this 1/f noise is not present (see e.g. the review of van der Klis 1995). Ebisawa (1991) found in a systematic study of Ginga observations of 6 BHC that the time scales of variability for the soft and hard components are often different. The soft component is usually roughly stable on time scales of 1 day or less, while the hard component exhibits large variations down to msec time scales. If the time scales of BHC and quasars scale with the mass of the compact object the above two time scales translate to $10^4$ years and 0.1 day respectively for our quasar sample. For a sample of quasars with similar redshifts and luminosities this predicts a rather small scatter in the soft component and a bigger scatter in the hard component. ROSAT results go in this direction. Laor et al. (1994, 1997) find that (for their sample of 23 low--z PG quasars) the scatter in the normalized 2 keV luminosity is significantly larger than that in the 0.3 keV luminosity. We conclude that the analogy between AGN and Galactic BHC seems to hold qualitatively for their X-ray variability properties. An alternative and intriguing possibility to explain the correlation between X-ray variability amplitude and spectral shape is that a component generated closer to the black hole dominates the emission of steep $\\alpha_X$ quasars, as in the spherically converging optically thick flow proposed by Chakrabarti and Titarchuk (1995) to characterize BHC in the high and soft state. If this is the case then we would again expect that the spectrum of the steep $\\alpha_X$ quasars remain steep above 2 keV (and up to $m_ec^2$ according to Chakrabarti and Titarchuk). Observations with the ASCA and BeppoSAX satellites instruments, which are sensitive up to 10 keV, (Brandt et al 1997, Fiore et al 1998b, Comastri et al 1998) suggest that this is indeed the case. Schwartz and Tucker (1988) suggested that AGN emission can be produced by an ensemble of acceleration sites (i.e. shock waves) with different electron spectral indices and therefore emitting power laws with different photon indices. In this picture the spectrum is a quadratic function in log E and the mean index at a given energy arises from the greatest number of individual acceleration sites. Variability would be greatest for both steepest energy index and flattest energy index sources, each of which are dominated by fewer individual regions. This is contradicted by the present observations, unless the Soft X-ray flat quasars become still flatter (and more variable) above 2 keV. High energy X-ray spectroscopy and variability studies are again needed to obtain a definitive answer. A RossiXTE monitoring campaign of 4 PG quasars, with a sampling similar to that used in the HRI campaign, is in progress and could allow us to clarify this point." }, "9803/astro-ph9803295_arXiv.txt": { "abstract": "We have searched for molecular gas towards the nucleus of four galaxies known to harbor a water vapor megamaser. CO(1\\raw0) emission of NGC 2639 and NGC 5506 was strong enough to allow us to map their inner regions. Weak emission from Mrk 1210 was detected and Mrk 1 was not detected at all. We report the tentative detection of the CO(2\\raw1) line in NGC 5506. After this work, 12 of the 18 known galaxies harboring a water vapor megamaser have been observed in CO. The molecular gas content in the inner regions of water megamaser galaxies ranges from 5$\\times$$10^7$ to 6$\\times$$10^9~\\Msun$. The circumnuclear molecular gas surface density also extends over nearly two orders of magnitude. The maser luminosity is correlated neither with the total amount of molecular gas found in the inner few kpc of these galaxies nor with global properties of the molecular gas such as surface density or filling factor; it is also independent of the infrared and optical luminosities. The only significant correlation we have found involves the maser luminosity and the low frequency radio continuum flux density. We conclude that the maser activity is intrinsically related to the energy of the active galactic nucleus whereas the intensity and even the presence of a water megamaser is independent of the molecular gas global properties such as the molecular gas content and surface density in the inner galactic regions. We have also found a pos\\-sible anti\\-cor\\-re\\-lation between the molecular gas surface density and the rate of the megamaser variations. A higher molecular gas abundance in the inner region could lead to higher maser variability because of larger nuclear flux variations due to the more variable gas infall, and/or because of more frequent interactions of the pumping agent with molecular gas condensations. ", "introduction": "The first water megamaser was discovered towards the nucleus of NGC 4945 (dos Santos \\& Lepine 1979). Strong emission at 22 GHz was detected, with a luminosity about 100 times higher than that of W49, the most powerful galactic water maser. Until 1985 four additional water megamasers were discovered, towards the nuclei of the Circinus Galaxy, NGC 1068, NGC 4258 and NGC 3079. During the following decade no more extragalactic water megamasers were found. In 1994 a survey towards active galactic nuclei was carried out by Braatz et al. (1994), leading to the discovery of five new water megamasers, doubling the number known to that date. Recently, Braatz et al. (1996) have reported the discovery of 6 additional water megamasers towards active galactic nuclei. The last ones up to now have been found in NGC 5793 (Hagiwara et al. 1997) and in NGC 3735 (Greenhill et al. 1997b). \\begin{table*}[t] \\begin{center} \\caption{Adopted parameters for the observed water megamaser galaxies} \\begin{tabular}{lcccc} \\hline Parameter & NGC 2639 & NGC 5506 & Mrk 1 & Mrk 1210 \\\\ \\hline \\hline $\\alpha_{2000}$$^a$ & \\hms[8 43 38.0] &\\hms[14 13 14.9] &\\hms [1 16 7.25] & \\hms[8 4 6.0 ] \\\\ $\\delta_{2000}$$^a$ & \\gms[50 12 20 3] &\\gms[$-$3 12 26 7] &\\gms[33 5 22 2] & \\gms[5 6 50 4] \\\\ Morphological type$^a$ & SA(r)a & SA pec sp & S & S \\\\ Activity$^a$ & LINER & Sy 2 &Sy 2 &Sy 2 \\\\ Heliocentric velocity (\\kms)$^a$ & 3236 &1816 & 4824 & 4043 \\\\ \\VLSR (\\kms) & 3235 &1825 & 4825 & 4030 \\\\ Distance (Mpc)$^b$ & 44 & 24 &64 &54 \\\\ Position angle (deg)$^c$ & 140 & 91 & --- & ---\\\\ Inclination (deg) & 58 & 80 & 56 & 3 \\\\ $D_{25}$ (arcmin)$^c$ & 1.8 & 2.8 & 0.8 & 0.8 \\\\ Linear scale (pc arcsec$^{-1}$) &209 &118 &312 & 260\\\\ $\\log L_{\\rm IR}$ (\\Lsun)$^b$& 10.21 &10.35 & --- & 10.58\\\\ $\\log L_{\\rm FIR}$ (\\Lsun)$^b$& 9.95 & 9.85 & ---& 9.85 \\\\ \\hline \\multicolumn{5}{l}{$^a$ Obtained from the NASA Extragalactic Database (NED)}\\\\ \\multicolumn{5}{l}{$^b$ See more information in the text (Sect. 2)}\\\\ \\multicolumn{5}{l}{$^c$ Obtained from the RC3 catalog (de Vaucouleurs et al. 1991)}\\\\ \\end{tabular} \\end{center} \\end{table*} Water megamasers, unlike the galactic and normal extragalactic masers, have been detected at the nuclei of distant galaxies. Interferometric observations (Claussen \\& Lo 1986; Greenhill et al. 1995a; Greenhill et al. 1996) have shown that megamaser emission comes from within a region around the galactic nucleus whose radius is in general smaller than 1 pc. Another important point is that all the galaxies harboring a water megamaser present some level of activity. This fact has led to suppose that the maser emission mechanism is identical to galactic masers. The huge difference in the energy involved is explained in terms of the pumping source: the energy source of the megamasers is the central object of the active nucleus, a much more powerful source than the central star powering the galactic masers. Braatz et al. (1997) have recently examined the conditions for detectability of water megamasers in terms of a variety of properties of the active galaxies, but not including the molecular gas content. It is known (Heckman et al. 1989) that Seyfert 2 have abnormally large quantities of dust and gas if compared with Seyfert 1 galaxies. On the other hand LINER galaxies are thought to be simply extensions of Seyfert 2 galaxies to lower luminosities, photoionized by a weaker AGN spectrum (Osterbrock et al. 1993). Thus, the fact that no water megamasers have been found in Seyfert 1 nuclei, while all of them are in either Seyfert 2 or LINER galaxies seems to indicate that a high concentration of molecular gas in the inner regions of the galaxies is a key parameter for the megamaser emission to be produced. To check this hypothesis and to find if the properties of the masers are related to those of the molecular gas, we have carried out a study of the molecular gas content and its properties in several of the water vapor megamasers for which no previous data existed or were incomplete. The megamasers we observed are four of those discovered by Braatz el al. in 1994: NGC 2639, NGC 5506, Mrk 1 and Mrk 1210. In our analysis we have used the available CO data for other water megamaser galaxies found in the literature (Heckman et al. 1989; Planesas et al. 1989; Sahai et al. 1990; Aalto et al. 1991; Wang et al. 1992 and Young et al. 1995). \\begin{figure*}[t] \\vspace{6cm} \\special{psfile=ms7133f1.ps hoffset=-40 voffset=-456} \\caption{CO(1\\raw0) emission profiles towards the central region ($23\\as$ diameter) of the three detected water megamaser galaxies. Velocity resolution is $21 \\kms$. The two NGC objects show double peak spectra, which may indicate the existence of a molecular gas ring surrounding the central AGN at a kpc scale. The spectrum of Mrk 1210 is single-peaked, a possible indication that the expected molecular gas ring is seen face-on. This spectrum was obtained in May 1997} \\end{figure*} ", "conclusions": "We have searched for molecular gas towards the nucleus of four galaxies known to harbor a water vapor megamaser, and detected the CO(1\\raw0) emission in three of them and the CO(2\\raw1) emission in one. With this work 12 of the 18 known water megamaser galaxies have been observed in CO, and only the most distant of the observed ones, Mrk 1, has not been detected yet. \\begin{enumerate} \\item The 12 water megamaser galaxies with molecular gas data available are not an homogeneous set regarding their molecular gas properties. The amount of H$_2$ in their circumnuclear regions ranges from 5$\\times$$ 10^7$ to 6$\\times$$ 10^9\\ \\Msun$. The extreme values of the H$_2$ surface density, $\\Sigma_{\\rm H_2}$, for the central kpc are 6 and 280 \\Msun\\ pc$^{-2}$. This parameter extends over a range of 2 orders of magnitude, a range similar to that of Seyfert galaxies, starburst galaxies or luminous infrared galaxies. The maser luminosity, $L_{\\rm maser}$, is not correlated to the total molecular gas mass. Therefore it seems that the total amount of molecular gas in the inner few kpc is not a fundamental parameter on which depends the existence and the average intensity of the water megamaser. \\item $L_{\\rm maser}$ is not correlated with $\\Sigma_{\\rm H_2}$ or to the filling factor of giant molecular clouds. Apparently the maser luminosity does not depend on the content of molecular gas in the inner kpc. However, the accumulation of clouds of dense molecular gas is believed to be necessary for the generation of a water megamaser. Observations with higher angular resolution of the molecular gas in the inner regions would help to solve the issue. \\item The only correlation we have found involving the maser emission and molecular gas parameters is between the rate of relative variation of the maser intensity and $\\Sigma_{\\rm H_2}$. This fact may indicate that a high abundance of molecular gas in the inner regions could lead to higher variability in the maser emission, on the one hand, due to the higher variability of the central pumping source produced by wider variations in the gas infall; on the other hand, due to the more frequent interactions of the pumping agent with molecular gas condensations. \\item $L_{\\rm maser}$ is not correlated to any other luminosity (infrared, optical, X-ray, blue). However, we have found some correlation between $L_{\\rm maser}$ and the global radio continuum flux density at 1.4 and 8.4 GHz. This fact supports the idea that $L_{\\rm maser}$ is a property related to the galactic nucleus (characterized by the radio luminosities) rather than to the inner galactic regions (characterized by the infrared luminosity and the molecular gas content). \\item Mrk 1210 stands as a peculiar object, having the highest star formation efficiency among water megamaser galaxies is spite of its relatively low molecular gas content. \\end{enumerate}" }, "9803/astro-ph9803156_arXiv.txt": { "abstract": "We study the FIR and UV-visible properties of star forming galaxies in the nearby Universe. This comparison is performed using the local luminosity functions at UV and FIR wavelengths and on individual starburst galaxies for which photometric data from UV to NIR and FIR are available. The FIR and UV luminosity functions have quite different shapes~: the UV function exhibits a strong increase for low luminosity galaxies whereas the FIR tail towards ultra luminous galaxies ($\\rm L > 10^{11} L\\odot$) is not detected in UV. The comparison of the FIR and UV local luminosity densities argues for a rather moderate extinction in nearby disk galaxies. The galaxies selected to be detected in FIR and UV are found to be located in the medium range of both luminosity functions. An emphasis is made on starburst galaxies. For a sample of 22 of these objects, it is found that the UV (912-3650 $\\rm \\AA$), the visible (3600-12500 $\\rm \\AA$) and the NIR (12500-22000 $\\rm \\AA$) wavelength range contribute $\\sim 30\\%$, $\\sim 50\\%$ and $\\sim 20\\%$ respectively to the total emerging stellar emission (for a subsample of 12 galaxies for the NIR and visible light). The mean ratio of the dust to bolometric luminosity of these galaxies is 0.37$\\pm$0.22 similar to the ratio found for normal spiral galaxies. Only 4 out of the 22 galaxies exhibit a very large extinction with more than 60$\\%$ of their energy emitted in the FIR-submm range. The mean extinction at 2000$\\rm \\AA$ is found to be $\\sim 1.2$ mag although with a large dispersion. The UV, visible and NIR emissions of our sample galaxies are consistent with a burst lasting over $\\sim 1$ Gyr. The conversion factor of the stellar emission into dust emission is found to correlate with the luminosity of the galaxies, brighter galaxies having a higher conversion factor. Since our sample appears to be representative of the mean properties of the galaxy population in FIR and UV, a very large conversion of the stellar light into dust emission can no longer be assumed as a general property of starburst galaxies at least in the local Universe. Instead a larger amount of energy emerging from the present starburst galaxies seems to come from the stars rather than from the dust. We compare the UV properties of our local starburst galaxies to those of recently detected high redshift galaxies. The larger extinction found in the distant galaxies is consistent with the trend we find for the nearby starburst galaxies namely the brighter the galaxies the lower the escape fraction of stellar light. \\keywords { Galaxies: starburst --Galaxies: stellar content -- Infrared: galaxies-- Ultraviolet: galaxies -- dust, extinction} ", "introduction": "One of the most important challenge of modern astronomy is the detection of young primeval galaxies. Indeed, very significant progress has been made with the detection of very high redshift galaxies either from ground based observations (Steidel et al. 1996b) or in the Hubble Deep Field (e.g. Steidel et al. 1996a, Lowenthal et al. 1997). In order to understand the properties of high redshift galaxies and study the cosmological evolution of star-forming galaxies, it is crucial to properly characterize the properties of starburst galaxies in the local Universe. These nearby galaxies are forming stars with a very high rate and it it actually important to analyse their emission over the entire spectral range (from UV to FIR-submm) to study the efficiency of the dust extinction and know what is the spectral range (UV, visible, NIR or FIR) where most of their energy is emitted. If high redshift star forming galaxies are similar to their low-z counterparts studying the latter will bring some clues to detect the former. We can wonder whether the observation of the rest-frame UV continuum is the best way to detect high redshift galaxies or if the high obscuration from dust makes them emit more energy in the FIR. IRAS discovered infrared bright galaxies with prodigious star formation rates, a very high extinction and therefore a low optical flux (Sanders \\& Mirabel 1996). The most luminous FIR galaxies are produced by strong interactions or merging of molecular gas-rich galaxies which induce enormous starbursts. Such objects might be the progenitors of elliptical galaxies (Kormendy \\& Sanders 1992). In a \"bottom-up\" scenario of galaxy formation, numerous starbursts induced by merging are expected and the bulk of their emission would be in the FIR redshifted in the submm (e.g. van der Werf \\& Israel 1996). Mazzei et al. (1994) predict that more than 90 \\% of the energy emitted by a starburst is in the FIR range during the first Gyr. This percentage rapidly drops to reach $\\sim 30\\%$ for $\\sim 5$ Gyr- old objects. Models aimed at explaining the galaxy counts in optical and FIR predict that during intense phases of star formation the quasi totality of the stellar light is absorbed and re-radiated in the FIR wavelength range (Franceschini et al. 1994, Pearson \\& Rowan-Robinson 1996). Conversely, considerable effort has been carried out in the UV-optical study of star forming galaxies for some years. Calzetti, Kinney and co-workers have extensively used the IUE spectra of star-forming galaxies complemented with optical and IR data to characterize the star formation history and the extinction occurring in the central regions of these objects (Calzetti et al. 1994, Calzetti et al. 1995, Calzetti 1997a). Meurer et al. (1995) have used Faint Object Camera on board the {\\it Hubble Space Telescope} HST-FOC observations to study the morphology of some starburst galaxies. From these studies, a foreground distribution of dust and a rather grey extinction curve seems to be able to explain the spectral distribution of the central regions of starburst galaxies. The extinction found by Meurer et al. (1995) for nearby starburst galaxies is rather low: at 2000 $\\rm \\AA$ it lies between 0.08 and 1.9 mag (excluding NGC7552 at 3.13 mag). Nevertheless these studies deal with the central parts of starburst galaxies and may well not be valid for the global emission of these objects at least when longer wavelengths than UV are concerned (Buat et al. 1997). Analyses of the UV-optical and FIR global emissions of nearby spiral and irregular galaxies selected to be observed both in UV and in FIR led to a rather low extinction (Xu \\& Buat 1995, Buat \\& Xu 1996, Wang \\& Heckman 1996). An important result of these studies is that the UV non-ionizing stellar emission is likely to be the major cause of dust heating. The contribution of OB stars to the dust heating is estimated to amount to about 20$\\%$ of the total FIR emission in the Milky Way (Cox \\& Mezger 1989), almost the same contribution of the ionizing radiation to the dust heating is found for spiral galaxies (Xu \\& Buat, 1995) and for starburst objects (Calzetti et al. 1995). The comparison between the FIR emission (dust re-radiation) and the UV and optical emission (escaped stellar light) constrains the extinction. As a consequence the FIR to UV continuum ratio is a powerful indicator of the extinction occurring in galaxies (Meurer et al. 1995, Buat \\& Xu 1996, Wang \\& Heckman 1996). Recently, HST imaging of very high redshift galaxies complemented when possible by spectroscopic observations with the Keck Telescope have led to the discovery at high redshift (z$\\sim 2-3$) of compact star forming galaxies with a moderate size and a strong rest-frame UV emission (Steidel et al. 1996a) with sometimes more diffuse extended structures Lowenthal et al. 1997). Depending on the intrinsic UV spectrum adopted, the average extinction estimated at 1600 $\\rm \\AA$ from the rest-frame UV spectral energy distribution of these galaxies is of the order of 1.7 to 3 mag (Meurer et al. 1997, Calzetti 1997b). These significant average extinctions are therefore larger than the values estimated for nearby starburst galaxies. However, it must be noted that these high redshift galaxies are very luminous ($\\rm M_B < -21$) when compared to the mean luminosity of nearby starburst galaxies studied by Meurer et al. (1995) ($\\rm = -18.6$) and the extinction is known to correlate with the luminosity of galaxies (Giovanelli et al. 1995, Wang \\& Heckman 1997). Moreover, the selection biases are very strong towards very luminous galaxies with a strong UV continuum and it cannot be excluded that high redshift galaxies almost entirely hidden by the dust are missing from these observations in the rest-frame UV (Mobasher et al. 1996, Burigana et al. 1997). The selection biases in the recent detections of high z galaxies are difficult or even impossible to quantify in the absence of similar observations at other wavelengths corresponding to longer than UV rest frame emissions (NIR or FIR). A first step is to estimate the importance of such a bias in the local Universe. Such a study is also crucial to compare the properties of high redshift galaxies to those in the nearby Universe. At this aim we will adopt a global approach to study the local Universe which consists in comparing the luminosity functions and the luminosity densities in UV and FIR. The UV wavelength range is particularly interesting since it is a tracer of the recent star formation rate as already mentioned and observations in the visible range of high z ($>2$) galaxies correspond to their UV rest frame emission. More specifically we will compare the amount of energy locked up in FIR to the amount of energy directly emitted in UV in the local Universe. The comparison of these global values with individual galaxies selected to be observed both in UV and FIR will allow to discuss how such samples of individual galaxies are representative of the mean properties of the local Universe. We will also investigate the specific case of a sub sample of nearby starburst galaxies detected in UV, visible, NIR and FIR in order to compare their global dust and stellar emission and to estimate what fraction of the emission of stars is converted into dust emission as well as the relative contribution of the UV, visible and NIR spectral ranges to the observed stellar emission. Such estimates will lead to predict what spectral range is more favorable for the detection of high redshift starbursts under the hypothesis that they are similar to their nearby counterparts. The main limitation to this approach is that we deal with global fluxes integrated over the galaxies whereas the starburst often occurs in the central parts. Moreover the galaxies at high redshift so far detected seem to have compact morphologies. Nevertheless as it will be shown in section 4 a large fraction of the UV emission of a starburst galaxy is likely to come from the starburst itself making valid a study on the global fluxes as soon as this wavelength is concerned. It is also the case for the FIR emission since the UV (ionizing and non ionizing) emissions is the major contributor to the dust heating), especially in starbursting objects. Obviously, more care must be taken when dealing with the visible and NIR emission: at these wavelengths the contribution of the underlying old stellar population present in local starburst galaxies is very large even dominant. Endly, available FIR fluxes on large samples are integrated over the galaxies due to the poor resolution of the IRAS satellite and dealing with global fluxes allows a reliable comparison of the emission of galaxies at different wavelengths. Beyond the detection of high redshift star forming galaxies it is necessary to estimate a quantitative star formation rate (SFR) for these galaxies. The deduction of such a quantity from the observed rest-frame UV continuum relies almost entirely on the amount of the extinction with only a moderate dependence on the star formation history (e.g. Meurer et al. 1997, Calzetti 1997). More specifically after $\\sim 5~10^7$ years of constant star formation rate the UV flux (912-3650 $\\rm \\AA$) reaches 80 $\\%$ of its stationnary value calculated for $\\rm 10^{10}$ years of constant star formation (from Bruzual \\& Charlot 1993). From a global energetic budget, we will try to constrain the amount of extinction and bring some clues to this difficult problem. ", "conclusions": "\\subsection{ The local Universe} We have considered galaxies selected to be detected photometrically both in UV and FIR. An analysis of the local luminosity functions at both wavelengths shows that the galaxies selected in this way have a FIR and a UV emission representative of the mean population of galaxies in the local Universe. From a sample of 22 starburst galaxies, the mean escape fraction of the stellar light (ratio of the stellar luminosity to the bolometric one) is found to be $63\\%\\pm 22\\%$, i.e. similar to what is found for normal galaxies. This escape fraction exhibits a strong decrease with increasing galaxy luminosity ranging from $\\sim 80\\%$ at faintest B magnitudes ($\\rm M_B~>-17$) to $\\sim 10\\%$ for $\\rm M_B~\\sim~ -20$. This result is quite different to that of Pearson \\& Rowan-Robinson (1996) who found that the escape fraction of the stellar light in starburst galaxies cannot exceed 5-10$\\%$ from an analysis of deep counts in visible and NIR and assuming a very strong cosmological evolution for the starburst population. Such a low fraction of escaped light is consistent with the studies of FIR bright galaxies but seems not to be a generic property of starbursting objects, at least in the local Universe. \\subsection {Comparison with high redshift galaxies} We can compare our results found for nearby starbursts to high redshift galaxies recently detected. A limitation to this comparison might be that we deal with integrated fluxes on nearby galaxies which contain an underlying stellar population pre-existing to the starburst. To our knowledge, there is no evidence for the presence or not of such an older population in high redshift starburst galaxies (z $\\sim$ 3). Nevertheless, as discussed in the paper, the UV range is largely dominated by the emission of the newly formed stars and the comparison of the high redshift galaxies to present-day starburst ones in this wavelength range is justified. From an analysis of the slope of the rest-frame UV continuum, an extinction has been estimated for high z galaxies (Meurer et al. 1997, Calzetti 1997b, Pettini et al. 1997). A mean extinction of 1.65 mag at 1650 $\\rm \\AA$ is found by Calzetti for a star formation rate constant over $\\sim 1$ Gyr which implies an unreddened slope for the UV continuum $\\rm \\beta=-2$. This value is in agreement with Pettini et al.'s estimates. Meurer et al. have adopted a different star formation law with a starburst lasting 10 $\\rm Myr$, leading to a larger proportion of very young stars and a steeper UV slope ($\\rm \\beta=-2.5$). Such a very extreme scenario gives an extinction as large as 3 mag. at 1600 $\\rm \\AA$ which can be considered as a very upper limit. The extinction found for nearby starburst galaxies is lower than the value found by Calzetti or Pettini et al. at high redshift. But the galaxies observed at high redshift are much more luminous that the nearby starburst galaxies studied by us or by Meurer et al. (1995) (Meurer et al. 1997, Lowenthal et al. 1997, Pettini et al. 1997). As an example, an extinction of 1.6 mag at 1600 $\\rm \\AA$ gives $\\rm F_{dust}~/~F_{bol} = 0.6$ and $\\rm F_{dust}~/~F_{UV} = 4.5$ using the extinction curve of Calzetti et al. (1994), a foreground screen and a star formation rate constant over 1 Gyr. Extrapolating Fig.~5, these ratios correspond to the brightest galaxies with $\\rm M_B\\le -20$. If the trends found in Fig.5 are mainly due to the extinction, the extinction estimated in high redshift star forming galaxies is in agreement with the values found for nearby starburst galaxies when the galaxy luminosity is accounted for. We must point out, however, that low luminosity galaxies ( $\\rm M_B \\sim -16$) would be detectable with a NGST-type telescope in a reasonable amount of time as far as $\\rm z\\sim 9$ (if they exist). In a hierarchical scenario for the formation of the galaxies, it is likely that such galaxies would outnumber the large luminous galaxies presently detected (Ellis 1997). Obviously the knowledge of the local Universe properties in terms of the extinction occurring in galaxies as well as their spectral energy distribution, especially in the UV range, is essential to interpret the observations of the Universe at high redshift. It is also crucial to predict the best way to detect young galaxies during a phase of intense star formation. We show that a comparison of the FIR and UV emissions in nearby galaxies can bring some clues since these two emissions are very sensitive to the current star formation rate and to the extinction. A more straightforward method would be a systematic comparison of deep fields in UV and FIR. Such a preliminary comparison was already performed (Buat \\& Xu 1996) using FAUST (Deharveng et al. 1994) and IRAS observations on two 7.6$\\rm ^o$-wide fields in the central region of the Virgo cluster. Although the UV observations of FAUST were not very deep (limiting flux $\\rm 10^{14} - 10^{15}~ erg/cm^2/s/\\AA$ at 1600 $\\rm \\AA$), we have searched for galaxies detected by IRAS without any UV detected counterpart. Given the low sensitivity of the FAUST experiment, the lower limits obtained for the ratio of the FIR to UV luminosity are within the range of values found for the galaxies detected at both wavelengths (fig. 4 in Buat \\& Xu 1996). This is in agreement with the results found in this paper but needs to be confirmed with deeper observations. The UV observations carried out with the large field ($\\rm \\sim 2^o$) FOCA telescope (e.g. Donas et al. 1990) reach the magnitude 18 at 2000$\\rm \\AA$. Unfortunately up to now the FIR observations (made by the IRAS satellite) are not deep enough to be compared to these UV observations and we have to wait for new FIR large field instruments like WIRE to perform such a comparison." }, "9803/astro-ph9803010_arXiv.txt": { "abstract": "Metal line ratios in a sample of 13 quasar spectra obtained with the HIRES spectrograph on the KeckI telescope have been analyzed to characterize the evolution of the metagalactic ionzing flux near a redshift of 3. The evolution of \\ion{Si}{4}/\\ion{C}{4} has been determined using three different techniques: using total column densities of absorption line complexes, as in Songaila \\& Cowie\\markcite{sc96} (1996); using the column densities of individual Voigt profile components within complexes; and using direct optical depth ratios. All three methods show that \\ion{Si}{4}/\\ion{C}{4} changes abruptly at $z \\sim 3$, requiring a jump in value of about a factor of 3.4, and indicating a significant change in the ionizing spectrum that occurs rapidly between $z = 2.9$\\ and $z = 3$, just above the redshift at which Reimers et al.\\markcite{reim} (1997) detected patchy \\ion{He}{2} ${\\rm Ly}\\alpha$\\ absorption. At lower redshifts, the ionization balance is consistent with a pure power law ionizing spectrum but at higher redshifts the spectrum must be very soft, with a large break at the He$^+$\\ edge. An optical depth ratio technique is used to measure the abundances of ions whose transitions lie within the forest and \\ion{C}{3}, \\ion{Si}{3} and \\ion{O}{6} are detected in this way. The presence of a significant amount of \\ion{O}{6} at $z > 3$\\ suggests either a considerable volume of \\ion{He}{3} bubbles embedded in the more general region where the ionizing flux is heavily broken, or the addition of collisional ionization to the simple photoionization models. ", "introduction": "\\label{intro} Whereas we know from the absence of any significant Gunn-Peterson effect even in the highest redshift quasars that hydrogen reionization of the intergalactic gas must have taken place at $z > 5$, we have much less information about the period at which the bulk of singly ionized helium converted to doubly ionized helium. Since late He$^+$\\ ionization may significantly change the temperature of the intergalactic gas it is critical to understand this heating if we are to correctly model the growth of structure in the IGM, and determine the mapping of the baryon density to observable quantities such as observations of the neutral hydrogen ${\\rm Ly}\\alpha$\\ forest and the He$^+$\\ ${\\rm Ly}\\alpha$\\ opacity. Phenomenological modelling of this event depends critically on the softness of the composite spectrum of the ionizing sources (e.g.\\ Miralda-Escud\\'e \\& Rees\\markcite{mr93} 1993; Madau \\& Meiksin\\markcite{mm} 1994), amplified by the subsequent radiative transfer, and such models cannot be considered reliable in predicting the high energy ($E > 54~{\\rm eV}$) metagalactic ionizing spectrum above the He$^+$\\ ionization edge, which determines the fraction of singly ionized helium. Our most direct information on the He$^+$\\ opacity is through observations of quasars whose spectra extend to the He$^+$\\ ${\\rm Ly}\\alpha$\\ wavelength. Despite the extreme difficulty of these measurements, successful observations of the He$^+$\\ ${\\rm Ly}\\alpha$\\ absorption have been made toward $z > 2.8$\\ quasars with HST (Jakobsen et al.\\markcite{jak} 1994; Hogan et al.\\markcite{hog} 1997; Reimers et al.\\markcite{reim} 1997) and of the $z = 2.72$\\ quasar HS1700+6416 with HUT (Davidsen et al.\\markcite{dkz} 1996; Zheng et al.\\markcite{zdk} 1998). The He$^+$\\ ${\\rm Ly}\\alpha$\\ opacity shows a marked decrease from a value of $\\tau = 3.2^{+\\infty}_{-1.1}$\\ in Q0302$-$003 at $z = 3.29$\\ to $\\tau = 1.0 \\pm 0.07$\\ in HS1700+6416. More remarkably, the Reimers et al.\\markcite{reim} observation of the intermediate redshift quasar HE2347$-$4342 at $z = 2.89$\\ shows both `troughs' and `voids' in the \\ion{He}{2} ${\\rm Ly}\\alpha$\\ observations, suggesting that at $z \\sim 2.8$\\ we are seeing fully ionized \\ion{He}{3} bubbles interspersed among as yet unionized He$^+$\\ regions and that it is at this point that the porosity of \\ion{He}{3} regions is approaching unity. The recent discovery that the bulk of ${\\rm Ly}\\alpha$\\ forest absorption with $N({\\rm H~I}) > 3\\times 10^{14}~\\rm cm^{-2}$\\ contains associated metals (Songaila \\& Cowie\\markcite{sc96} 1996, hereafter SC) gives us an alternative approach to the problem since ionization balance in these forest metals provides a diagnostic of the shape of the metagalactic flux in the neighborhood of the He$^+$\\ edge. As was first noted in Songaila et al.\\markcite{shc} (1995), the value of \\ion{Si}{4}/\\ion{C}{4} in high ionization systems is critically dependent on the He$^+$\\ ionization edge break strength. This has subsequently been investigated in more detail by SC, Savaglio et al.\\markcite{sav} (1997) and Giroux \\& Shull\\markcite{gs97} (1997), among others. As will be discussed further here, the bulk of the forest metal line systems at the currently observed redshifts ($z \\sim 2 - 4$) are high ionization (\\ion{C}{2}/\\ion{C}{4} $\\ll 0.1$) and so, unless there is a strong break at the edge, they will have \\ion{Si}{4}/\\ion{C}{4} $\\ll 0.1$\\ even for the higher Si/C abundances characteristic of low metallicity systems. Therefore, once the He$^+$\\ is fully ionized and the IGM becomes relatively transparent to the integrated quasar spectrum (e.g.\\ in the metagalactic spectrum of Haardt \\& Madau\\markcite{hm} 1996), the observed \\ion{Si}{4}/\\ion{C}{4} values should fall in this low range. SC and Savaglio et al.\\markcite{sav} (1997) have shown that, whereas this is generally true at $z < 3$, much higher \\ion{Si}{4}/\\ion{C}{4} values are regularly seen at $z > 3$, suggesting a significant change above this redshift, which would be consistent with the interpretation of the Reimers et al.\\markcite{reim} (1997) observations as showing the redshift at which \\ion{He}{3} bubbles begin to overlap. Boksenberg\\markcite{bok} (1998) has recently questioned this result, based on an analysis of the the redshift evolution of the ion ratios in the separate Voigt profile components in complex systems rather than in the integrated column densities of the complexes. However, his analysis is based on rather a small sample of systems without clear selection criteria. In this paper, I shall use the largest sample to date to demonstrate unambiguously that, irrespective of the method of analysis, there is indeed a rapid jump in the value of \\ion{SI}{4}/\\ion{C}{4} at a redshift just below 3, and that the ionization stages in the metals are consistent with this being, for most systems, the point at which they change from being ionized by a metalgalactic spectrum that is, at $z > 3$, heavily broken above 54~eV to one that is only mildly broken at the lower redshifts. The sample and the data reduction are described in \\S 2 and the reader who is primarily interested in the results could safely skip this section. The evolution of the \\ion{C}{2}/\\ion{C}{4} and \\ion{Si}{4}/\\ion{C}{4} values with redshift is described in \\S 3 where I show that the presence of a jump in \\ion{Si}{4}/\\ion{C}{4} values at $z$\\ just under 3 is highly significant and does not depend on the method of analysis, whether by total column densities of complex, by the column densities of individual Voigt components, or by directly analyzing the distribution of optical depth ratios. In \\S 4 I consider the overall ionization balance including intermediate ionization stages such as \\ion{C}{3} and \\ion{Si}{3} and high ions such as \\ion{N}{5} and \\ion{O}{6} using, for those lines that lie primarily within the forest, the optical depth distributions of the ensembles of such lines, which provides a new and robust technique for determining their properties. The overall ionization balance is broadly consistent with unbroken power law photoionization at $z < 3$; however, the observation of significant amounts of \\ion{O}{6} at $z > 3$\\ requires either the presence of a considerable volume of \\ion{He}{3} bubbles permeating regions where the ionizing flux is heavily broken, or the addition of collisional ionization to the simple photoionization models. Finally, the conclusions are briefly summarised in \\S 5. ", "conclusions": "\\label{conc} I summarise the results of the paper by noting that, irrespective of the analysis methodology adopted, there is a significant change in the ionization balance of forest metal lines which occurs just below a redshift of 3. At lower redshifts, the ionization balance in the forest lines is fully consistent with a pure power law ionization spectrum with an index of $-1.8$\\ but at higher redshifts the high values of \\ion{Si}{4}/\\ion{C}{4} seen in most of the forest clouds despite generally low \\ion{C}{2}/\\ion{C}{4} values implies that the ionizing flux must be very soft, with a large break at the He$^+$\\ edge. The change occurs quite rapidly between $z = 2.9$\\ and $z = 3$, just above the redshift at which highly patchy \\ion{He}{2} ${\\rm Ly}\\alpha$\\ absorption is seen in the quasar HE~2347$-$4342 (Reimers et al.\\markcite{reim} 1997). The simplest explanation seems to be that we are seeing the redshift at which \\ion{He}{2} ionizes completely to \\ion{He}{3} as the \\ion{He}{3} Str\\\"omgren spheres overlap." }, "9803/astro-ph9803226_arXiv.txt": { "abstract": "The 6.4~s X-ray pulsar \\src\\ was observed by \\sax\\ in 1997 May. This source belongs to the class of ``anomalous'' pulsars which have pulse periods in range 5--11~s, show no evidence of optical or radio counterparts, and exhibit long-term increases in pulse period. The phase-averaged 0.5--10 keV spectrum can be described by an absorbed power-law and blackbody model. The best-fit photon index is 2.5$\\pm$0.2 and the blackbody temperature and radius are 0.64$\\pm$0.01~keV and $0.59 \\pm 0.02$~km (for a distance of 3~kpc), respectively. The detection of blackbody emission from this source strengthens the similarity with two of the more well studied ``anomalous'' pulsars, 1E\\,2259+586 and 4U\\,0142+614. There is no evidence for any phase dependent spectral changes. The pulse period of $6.45026 \\pm 0.00001$~s implies that \\src\\ continues to spin-down, but at a slower rate than obtained from the previous measurements in 1994 and 1996. ", "introduction": "\\label{sec:introduction} \\src\\ is an unusual pulsar, with a period of 6.4~s and a soft spectrum, discovered during {\\it Einstein}\\/ observations of the Carina nebula (Seward et al.\\ 1983). The source has been spinning down for at least the last 17 years (Mereghetti 1995; Corbet \\& Mihara 1997). The spectrum has been modeled by an absorbed power-law with a photon index, $\\alpha$, of $\\sim$2--3 (Seward et al.\\ 1983; Corbet \\& Mihara 1997). This spectral shape is softer than those of typical high-luminosity X-ray pulsars. Despite a small error box, no optical counterpart has been identified, with a limiting magnitude of $m_{V}\\sim20$ which excludes the presence of a massive companion (Mereghetti et al.\\ 1992). A recent observation with the {\\it RossiXTE}\\/ satellite provides a strong upper limit to the projected X-ray semi-major axis of 0.06~lt-s for orbital periods between 200~s and $\\sim$1~day (Mereghetti et al.\\ 1997). The lack of an optical counterpart and orbital Doppler shifts argue against a binary model for \\src, unless the companion has a low mass. Mereghetti et al.\\ (1997) find that a probable upper limit to the mass of a Roche lobe filling main sequence companion is $\\sim$0.3 \\Msun. Masses up to $\\sim$0.8 \\Msun\\ are allowed in the case of a helium-burning companion filling its Roche lobe. The properties described above are similar to those of a small number of other X-ray pulsars with spin periods in the 5--11~s range (Mereghetti \\& Stella 1995), such as 1E\\,2259$+$586 and 4U\\,0142+614. These form a class of so-called ``anomalous'' pulsars, with clearly different properties from the majority of systems. Although accretion from a very low mass companion cannot be excluded, the lack of evidence for a binary nature from any of these systems has stimulated models where the X-ray emission originates from a compact object that is not in an interacting binary system. While an isolated, massive, white dwarf powered by the loss of rotational energy, as originally proposed for 1E\\,2259$+$586 (Paczynski 1990; Usov 1994), has been ruled out by the detection of a large increase in the spin-down rate of \\src\\ (Mereghetti 1995), other single object models, such as loss of magnetic energy of a strongly magnetized neutron star (Thompson \\& Duncan 1993), or an isolated neutron star accreting from a circumstellar disk (Corbet et al. 1995; van Paradijs et al. 1995), may be applicable. We present a study of \\src\\, based on data obtained with the \\sax\\ satellite. We focus on the X-ray spectrum at energies $<$10 keV and on the pulse period history. As with some of the other ``anomalous'' pulsars, we find evidence for the presence of a blackbody spectral component. Since this component is not observed from the majority of other accreting X-ray pulsars, this strengthens the similarity between \\src\\ and the other better studied ``anomalous'' pulsars. ", "conclusions": "The 0.5--10~keV spectrum of \\src\\ can be described by the sum of an absorbed power-law and blackbody models. The existence of a blackbody component in \\src\\ was first suggested by the ASCA results of Corbet \\& Mihara (1997). However, these authors are unable to clearly discriminate between this two component model, which gives a \\rchisq\\ of 0.94 for 332 dof, and an absorbed power-law which also provides an acceptable description of the ASCA data with a \\rchisq\\ of 1.02 for 337 dof. The \\sax\\ results reported here clearly require that the \\src\\ spectrum differs significantly from an absorbed power-law. This deviation in consistent with a blackbody, but we cannot exclude the possibility that it has another form. A discussion about the physical interpretation of the alternate models can be found in White et al.\\ (1996). This discussion is applicable to \\src, since its spectral shape is roughly similar to that of 4U\\,0142+614. The reason for the significant \\sax\\ detection of spectral complexity is probably related to the combination of good energy resolution and extended low energy coverage of the LECS, together with the longer \\sax\\ exposure. 1E\\,2259+586 (Corbet et al.\\ 1995) and 4U\\,0142+614 (White et al.\\ 1996) have also been successfully fit with absorbed power-law and blackbody spectral models. Since the majority of X-ray pulsar spectra are fit by absorbed power-law models in the 0.5--10~keV energy range (e.g., White et al. 1983), our results strengthen the similarity between \\src\\ and the other ``anomalous'' pulsars. Table \\ref{tab:bbprops} is a compilation of the spectral properties of these ``anomalous'' pulsars. The blackbody component in these systems has been interpreted as evidence for quasi-spherical accretion onto an isolated neutron star formed after common envelope evolution and spiral-in of a massive X-ray binary (White et al.\\ 1996). In this case, the accretion flow results from the remaining part of the massive star's envelope and may consist of two components (Ghosh et al. 1997). A low-angular momentum component gives rise to the blackbody emission from a considerable fraction of the neutron star surface, while a high-angular momentum component forms an accretion disk and is responsible for the power-law emission and the long term spin-down evolution. The area for the blackbody emitting surface obtained for \\src\\ ($\\sim$0.59 d$_{\\rm 3kpc}$ km) is smaller than those of the other ``anomalous'' systems. However, the distance to \\src\\ is poorly constrained and the assumed distance of 3~kpc could be considered as a lower limit since the measured ${\\rm N_H}$ implies that it lies behind the Carina Nebula at 2.8~kpc (Seward et al.\\ (1986). White et al.\\ (1996) propose that the low pulsed fraction observed from 4U\\,0142+614 results from the large polar cap area in this system. This is consistent with the small polar cap area and high pulsed fraction reported here for \\src\\, but not with 1E\\,2259+586 and 1RXS{\\thinspace}J170849.0$-$400910, which both have large radii and a moderate pulsed fraction (see Table \\ref{tab:bbprops}). Interestingly, the best-fit blackbody temperature of 0.64~keV is somewhat higher than for 1E\\,2259+586 and 4U\\,0142+614 which may be related to the smaller area of the polar caps. Summarizing, we find that for \\src\\ (i) the blackbody temperature is higher, (ii) the blackbody radius is smaller, (iii) the power-law index is smaller (i.e. the spectrum is harder), and (iv) the pulsed fraction is higher than for the other ``anomalous'' pulsars (with the exception of RX{\\thinspace}J0720.4$-$3125 where the blackbody parameters are obtained from a one component fit). The phase-resolved spectra are not well fit with a constant contribution from the blackbody component. This is unsurprising given the large pulse amplitude ($\\sim$70\\%), together with the large blackbody contribution (55\\%; 2--10~keV) to the phase averaged flux. The probable constancy of the pulsed fraction over the 1--10~keV energy range and the lack of any spectral dependence on pulse phase (see Sect.\\ \\ref{subsec:src_pulsetiming}) imply that the {\\it whole} spectrum is varying in a similar manner; i.e. the pulsed component cannot be attributed solely to either the blackbody or the power-law components. Furthermore, the long term variations in source flux (Fig.\\ \\ref{fig:fluxhistory}) and the approximate constancy of the pulsed fraction (Table~\\ref{tab:pdata}) during this time again implies that either the luminosities of the two components are correlated, or that the underlying spectral shape is a more complex ``single component'' that happens to mimic a power-law and a blackbody. The spin-down rate of \\src\\ obtained from ROSAT and ASCA observations (Mereghetti 1995; Corbet \\& Mihara 1997), showed an 80\\% increase with respect to the value of $\\sim5 \\times 10^{-4}$ s yr$^{-1}$ measured before 1988 with \\ginga\\ and EXOSAT. This increasing trend was further extended by the {\\it RossiXTE}\\/ observation of 1996 July ($P$ = $6.449769 \\pm 0.000004$ s, Mereghetti et al.\\ 1997), yielding a $\\dot{P}$ of 13$\\times10^{-4}$ s yr$^{-1}$ after the ASCA measurement. The pulse period obtained with \\sax\\ lies significantly below the linear extrapolation from the previous two measurements, indicating, for the first time in \\src, a decrease in the overall spin-down rate. Table \\ref{tab:pdata} and Fig.\\ \\ref{fig:fluxhistory} indicate that there is no clear correlation between the observed long term spin-down rate ($\\dot{P}$) and the 2--10 keV source flux. At first sight this argues against an accretion hypothesis, but since $\\dot{P}$ is a long-term average and the flux an instantaneous measurement this comparison may not be valid. A much better comparison would be between the average flux and $\\dot{P}$ during the same time interval. Unfortunately, due to the faintness of the source, no such comparison is available. In view of the uncertainties in the average flux a good measure of the time variability of the source on timescales of months to years would be useful." }, "9803/astro-ph9803082_arXiv.txt": { "abstract": "The EROS and MACHO collaborations have each published upper limits on the amount of planetary mass dark matter in the Galactic Halo obtained from gravitational microlensing searches. In this paper the two limits are combined to give a much stronger constraint on the abundance of low mass MACHOs. Specifically, objects with masses $10^{-7}~\\msun \\simlt m \\simlt 10^{-3}~\\msun$ make up less than $25$\\% of the halo dark matter for most models considered, and less than $10$\\% of a standard spherical halo is made of MACHOs in the $3.5\\ee{-7}~\\msun < m < 4.5\\ee{-5}~\\msun$ mass range. ", "introduction": " ", "conclusions": "" }, "9803/astro-ph9803177_arXiv.txt": { "abstract": "We present a simple method for determining the (correlated) uncertainties of the light element abundances expected from big bang nucleosynthesis, which avoids the need for lengthy Monte Carlo simulations. Our approach helps to clarify the role of the different nuclear reactions contributing to a particular elemental abundance and makes it easy to implement energy-independent changes in the measured reaction rates. As an application, we demonstrate how this method simplifies the statistical estimation of the nucleon-to-photon ratio through comparison of the standard BBN predictions with the observationally inferred abundances. ", "introduction": "Big bang nucleosynthesis is entering the precision era \\cite{Sc97}. On the one hand, there has been major progress in the observational determination of the abundances of the light elements D \\cite{Bu97,We97}, $^3$He \\cite{Ba94,Gl96}, $^4$He \\cite{Ol97a,Iz97}, and $^7$Li\\cite{Ry96,Bo97}, although the increasing precision has highlighted discrepancies between different measurements (see Refs.\\cite{Mo97,Ho97,Le97} for recent assessments). Secondly, we have a sound analytical understanding of the physical processes involved \\cite{Be89,Es91} and the standard BBN computer code \\cite{Wa73,Ka92} which incorporates this physics is robust and can be easily altered to accomodate changes in the input parameters, e.g.\\ nuclear reaction rates \\cite{Sm93}. The comparison of increasingly accurate observationally inferred and theoretical abundances will further constrain the values of fundamental parameters, such as the nucleon density parameter (see, e.g., Ref.\\cite{Ol97b}) or extra degrees of freedom related to possible new physics beyond the Standard Model (see, e.g., Ref.\\cite{Sa96}). It goes without saying that error evaluation represents an essential part of such comparisons. Because of the complex interplay between different nuclear reactions, it is not straightforward to assess the effect on a particular elemental yield of the uncertainties in the experimentally determined reaction rates. The authors of Ref.\\cite{Kr90} first employed Monte Carlo methods to sample the error distributions of the relevant reaction cross-sections which were then used as inputs to the standard BBN computer code. This enables well-defined confidence levels to be attached to the theoretically predicted abundances, e.g. the abundance range within which say 95\\% of the computed values fall correspond to 95\\% C.L. limits on the expected abundance. It was later realized that error correlations are also relevant, and can be estimated with the same technique \\cite{Ke94,KeKr}. The Monte Carlo (MC) approach has since become the standard tool for comparing theory and data \\cite{Sm93,Kr94,Co95,Ha95,Ol97c}. However, although it can include refinements such as asymmetric or temperature-dependent reaction rate uncertainties \\cite{Sm93}, it requires lengthy calculations which need to be repeated each time (any of) the input parameters are changed or updated. Since we may expect continued improvement in the determination of the relevant parameters, it is desirable to have a faster method for error evaluation and comparison with observations. In this work we propose a simple method for estimation of the BBN abundance uncertainties and their correlations which requires little computational effort. The method, based on linear error propagation, is described in Sec.~II. A concrete application is given in Sec.~III, where theory and observations are compared using simple $\\chi^2$ statistics to obtain the best-fit value of the nucleon-to-photon ratio. In Sec.~IV we study with this method the relative importance of different nuclear reactions in determining the synthesized abundances. Conclusions and perspectives for further work are presented in Sec.~V. ", "conclusions": "We have shown that a simple method based on linear error propagation allows us to quantify the uncertainties associated with the elemental abundances expected from big bang nucleosynthesis, in excellent agreement with the results obtained from Monte Carlo simulations. This method makes transparent which nuclear reaction rate is mainly responsible for the uncertainty in the abundance of a given element. If determinations of the primordial abundances improve to the point where the observational errors become smaller than the theoretical uncertainties (say for $^7$Li), this will enable attention to be focussed on the particular reaction rate whose value needs to be experimentally better known. We have also demonstrated that for standard BBN, our method enables the use of simple $\\chi^2$ statistics to obtain the best-fit value of $\\eta$ from the comparison of theory and observations. At present there are conflicting claims regarding the primordial abundances of, particularly, D and $^4$He, and different choices of input data sets imply values of $\\eta$ differing by a factor of $\\sim\\,3$. However this quantity can also be determined through measurements of the angular anisotropy of the cosmic microwave background (CMB) on small angular scales. Within a decade the forthcoming all-sky surveyors MAP and PLANCK are expected to pinpoint the nucleon density to within $\\sim5\\%$ \\cite{cmb}. Such measurements probe the acoustic oscillations of the coupled photon-matter plasma at the (re)combination epoch and will thus provide an independent check of BBN, assuming $\\eta$ did not change significantly between the two epochs.% \\footnote{New physics beyond the Standard Model can change $\\eta$, e.g. by increasing the photon number through massive particle decay \\cite{decay} or, more exotically, by {\\em decreasing} the photon number through photon mixing with a shadow sector \\cite{Ba91}. However such possibilities are strongly constrained by the absence of distortions in the Planck spectrum of the CMB \\cite{El92} and also, in the latter case, by the absence of Sakharov oscillations in the power spectrum of large-scale structure \\cite{Bi97}.} Nevertheless precise measurements of light element abundances, particularly $^4$He, are still crucial because they provide a unique probe of physical conditions, in particular the expansion rate at the BBN epoch. To illustrate, if $\\eta$ was determined by the CMB measurements to be $\\approx\\,2\\times10^{-10}$ (consistent with data set ``A''), but the abundance of $^4$He was established to be actually closer to its higher value of $\\approx\\,24\\%$ in data set ``B'', this would be a strong indication that the expansion rate during BBN was higher than in the standard case with $N_\\nu=3$ neutrinos. Although the number of $SU(2)$ doublet neutrinos is indeed 3, there are many light particles expected in extensions of the Standard Model, e.g. singlet neutrinos, which can speed up the expansion rate during nucleosynthesis \\cite{Sa96}. The generalization of our method to such non-standard cases is straightforward and we intend to present these results in a future publication \\cite{us}. It is clear that BBN analyses will continue to be important in this regard for both particle physics and cosmology." }, "9803/astro-ph9803207_arXiv.txt": { "abstract": "Hubble Space Telescope observations of the gravitational lens PG~1115+080 in the infrared show the known $z_l=0.310$ lens galaxy and reveal the $z_s=1.722$ quasar host galaxy. The main lens galaxy G is a nearly circular (ellipticity $\\epsilon < 0.07$) elliptical galaxy with a de Vaucouleurs profile and an effective radius of $R_e = 0\\farcs59\\pm0\\farcs06$ ($1.7 \\pm 0.2 h^{-1}$ kpc for $\\Omega_0=1$ and $h = H_0/100$ \\kmm). G is part of a group of galaxies that is a required component of all successful lens models. The new quasar and lens positions (3 milliarcsecond errors) yield constraints for these models that are statistically degenerate, but several conclusions are firmly established. (1) The principal lens galaxy is an elliptical galaxy with normal structural properties, lying close to the fundamental plane for its redshift. (2) The potential of the main lens galaxy is nearly round, even when not constrained by the small ellipticity of the light of this galaxy. (3) All models involving two mass distributions place the group component near the luminosity-weighted centroid of the brightest nearby group members. (4) All models predict a time delay ratio $r_{ABC}\\simeq 1.3$. (5) Our lens models predict $H_0=44\\pm4$ \\kmm\\ if the lens galaxy contains dark matter and has a flat rotation curve, and $H_0=65\\pm5$ \\kmm\\ if it has a constant mass-to-light ratio. (6) Any dark halo of the main lens galaxy must be truncated near $1\\farcs5$ ($4 h^{-1}$ kpc) before the inferred \\Ho\\ rises above $\\sim 60$ \\kmm. (7) The quasar host galaxy is lensed into an Einstein ring connecting the four quasar images, whose shape is reproduced by the models. Improved NICMOS imaging of the ring could be used to break the degeneracy of the lens models. ", "introduction": "Gravitational lens time delays offer a means of determining the Hubble constant that is purely geometrical and hence completely avoids the complications of the local distance scale (Refsdal 1964). The time delay for Q~0957+561 is now well-measured (Schild \\& Thomson 1997; Kundi\\'c et al. 1997; Haarsma et al. 1997), but significant systematic uncertainties remain due to the degeneracy between the mass of the primary lens galaxy and its host cluster (e.g. Grogin \\& Narayan 1996; Bernstein et al. 1997; Romanowsky \\& Kochanek 1998). No single lens is likely to be completely free of systematic uncertainties, so a reliable estimate of $H_0$ should rely on an ensemble of lenses. There are now three more systems with time delay estimates: PG~1115+080 (Schechter et al. 1997), B~1608+656 (Fassnacht et al. 1996), and B~0218+357 (Corbett et al. 1996), which need detailed exploration of their lens models to examine the systematic uncertainties. PG~1115+080 was the second gravitationally lensed quasar to be discovered (Weymann et al. 1980). The source is an optically selected, radio-quiet quasar at redshift $z_s=1.722$. Hege et al. (1981) first resolved the four quasar images (a close pair A1/A2, B and C), confirming the early model of Young et al. (1981) that the lens was a five-image system, one image being hidden in the core of the lens galaxy. Henry \\& Heasly (1986) detected the lens galaxy, followed by gradual improvements in the astrometry by Kristian et al. (1993; hereafter K93), and Courbin et al. (1997). The redshift of the lens galaxy was determined by Angonin-Willaime, Hammer \\& Rigaut (1993) and confirmed by Kundi\\'c et al. (1997) and Tonry (1998) to be $z_l=0.310$. Tonry also determined the central velocity dispersion of the lens galaxy: $\\sigma = 281\\pm25$ km s$^{-1}$. The spatial resolution of published data has always been insufficient to perform any surface photometry on the lens galaxy. Young et al. (1981) noted that the lens seemed to be part of a small group centered to the southwest of the lens, with a velocity dispersion of approximately $270\\pm70$ km s$^{-1}$ based on only four galaxy redshifts (Kundi\\'c et al. 1997). The group is an essential component of any model that successfully fits the lens constraints (Keeton, Kochanek \\& Seljak 1997; Schechter et al. 1997). Finally, Schechter et al. (1997) successfully determined two time delays between the images, which were reanalyzed by Barkana (1997) to give $\\Delta\\tau_{BC}=25.0^{+1.5}_{-1.7}$ days and the time delay ratio $r_{ABC}=\\Delta\\tau_{AC}/\\Delta\\tau_{BA}=1.13\\pm0.18$. These results were analyzed by Keeton \\& Kochanek (1997) and Courbin et al. (1997) to deduce $H_0=53_{-7}^{+15}$ km s$^{-1}$ Mpc$^{-1}$, with comparable contributions to the uncertainties from the time delay measurement and the models. The extreme variations are given in non-parametric form by Saha \\& Williams (1997), although some of these models may not be physical. We present new near-infrared observations of the PG~1115+080 system obtained with the Hubble Space Telescope (HST) NICMOS camera. These are the first results of the CfA-Arizona Space Telescope Lens Survey (CASTLES).\\footnote{A summary of gravitational lens data and model results, including CASTLES data, is available at the URL http://cfa-www.harvard.edu/castles.} After summarizing the observations in \\S2, we present improved astrometry in \\S2.1, the first surface photometry of the lens galaxy in \\S2.2, a discussion of lens models and the Hubble constant in \\S2.3, photometry of the nearby group in \\S2.4, and comments on the quasar host galaxy in \\S2.5. In \\S3 we comment on the strengths and limitations of this system as a cosmological tool. ", "conclusions": "New infrared data on PG~1115+080 affirms multiple-component gravitational lens systems as powerful cosmological tools. The major puzzle remaining in the PG~1115+080 system is the anomalous A1/A2 flux ratio. Our observations rule out differential extinction as an explanation, and microlensing is ruled out by its lack of variability. Since a flux ratio near 0.9 is a generic feature of the large scale potential near a fold caustic, only a potential perturbation intermediate between that produced by isolated stars (microlensing) and by the overall galaxy can explain the flux ratio. The potential of PG~1115+080 must be perturbed either by a satellite galaxy or a globular cluster. Mao \\& Schneider (1998) showed that such perturbations alter the time delay -- and so the inferred value of the Hubble constant -- fortunately by no more than 2-3\\%. Our improved astrometry greatly reduces some of the degeneracies in early models of the system. The group position is now well constrained and located near the luminosity centroid of the four bright group galaxies, and the lens galaxy is constrained to be nearly circular. Unfortunately, the degeneracies in the $H_0$ estimate have been exacerbated because with the revised astrometry the models no longer favor dark matter models over constant $M/L$ models for the main lens galaxy. Because all 4 images are located at nearly the same radial distance from the center of the lens galaxy, we do not expect the models to be sensitive to the radial mass profile of the lens galaxy (Kochanek 1991; Wambsganss \\& Pa\\'czynski 1994). We find \\Ho\\ ranging from $44\\pm4$ \\kmm\\ if the lens galaxy is modeled as a singular isothermal ellipsoid and the group as a singular isothermal sphere, to $65\\pm5$ ($72\\pm5$) \\kmm\\ for $\\Omega_0 = 1$ ($0.1$) if the lens galaxy has a constant $M/L$. Note that we find evidence for dark matter in our high value of $M/L$. A model with an adjustable truncation radius shows that the halo must be truncated on scales comparable to the ring diameter for $H_0$ to exceed $60$ \\kmm. Such a halo seems smaller than physically plausible given that the velocity dispersions of the group and the lens galaxy are comparable. Further progress in reducing the uncertainties depends on improving the time delay measurements and on making more detailed studies of the Einstein ring formed by the quasar host galaxy. First, for any given mass profile, most of the current errors in \\Ho\\ are due to the uncertainties in the time delays. Second, all the best fit models predict time delay ratios near $r_{ABC}=1.3$, consistent with the current measurement of $1.13 \\pm 0.18$. If nothing else, a more accurate measurement of the delay ratio than is now available would be a powerful test of the models. For any given lens mass profile, the ratio constraint would further reduce the parameter space for the position and mass of the group, or could be used to constrain more complicated models (e.g., Saha \\& Williams 1997). Deep new observations to determine the surface brightness of the ring accurately, combined with direct measurement of the point spread function at the time of the observations would probably permit us directly to break the degeneracy of the models. Only NICMOS, however, has the ability to make these difficult observations within a decade. No single technique or observation can tie down the Hubble constant -- the long history of unrecognized or underestimated systematic errors in this subject encourages humility. Nevertheless, PG~1115+080 demonstrates the potential of the gravitational lens approach. With recent reductions in age estimates of the oldest globular clusters (Chaboyer et al. 1998), and the likelihood that the mass density of the universe is lower than the Einstein-de Sitter case (e.g., Garnavich et al. 1998), the possibility of an age conflict in the big bang model has receded. The modeling of PG~1115+080 gives a plausible upper bound on the Hubble constant if we accept that the group is not a point mass and that the lens galaxy is unlikely to have a mass distribution that is more concentrated than its light distribution. This bound is $H_0 < 67$ ($72$) \\kmm\\ for $\\Omega_0 = 1$ ($0.1$). The most recent result of the HST Extragalactic Distance Scale Key Project is $H_0 = 72 \\pm5$ (random) $\\pm12$ (systematic) km s$^{-1}$ Mpc$^{-1}$ (Madore et al. 1998). Our upper limit is inconsistent with the upper end of the range from the Key Project, although it is consistent with the lower end of the range. There appears to be satisfactory concordance among the basic parameters of the big bang model, and between direct and indirect measures of the distance scale. Gravitational lenses can be expected to play an increasing role as versatile cosmological tools. \\vskip 1truecm" }, "9803/astro-ph9803031_arXiv.txt": { "abstract": "In this paper, we measure the ellipticities of 30 LSB dI galaxies and compare the ellipticity distribution with that of 80 dEs (Ryden \\& Terndrup 1994; Ryden et al.\\ 1998)\\markcite{ryden 1994; 1998} and 62 BCDs (Sung et al.\\ 1998).\\markcite{sung1998} We find that the ellipticity distribution of LSB dIs is very similar to that of BCDs, and marginally different from that of dEs. We then determine the distribution of intrinsic shapes of dI galaxies and compare to those of other type dwarf galaxies under various assumptions. First, we assume that LSB dIs are either all oblate or all prolate, and use non-parametric analysis to find the best-fitting distribution of intrinsic shapes. With this assumption, we find that the scarcity of nearly circular LSB dIs implies, at the 99\\% confidence level, that they cannot be a population of randomly oriented oblate or prolate objects. Next, we assume that dIs are triaxial, and use parametric analysis to find permissible distributions of intrinsic shapes. We find that if the intrinsic axis ratios, $\\beta$ and $\\gamma$, are distributed according to a Gaussian with means $\\beta_0$ and $\\gamma_0$ and a common standard deviation of $\\sigma$, the best-fitting set of parameters for LSB dIs is $(\\beta_0,\\gamma_0,\\sigma) = (0.66,0.50,0.15)$, and the best fit for BCDs is $(\\beta_0,\\gamma_0,\\sigma) = (0.66,0.55,0.16)$, while the best fit for dEs is $(\\beta_0,\\gamma_0,\\sigma) = (0.78,0.69,0.24)$. The dIs and BCDs thus have a very similar shape distribution, given this triaxial hypothesis, while the dEs peak at a somewhat more spherical shape. Our results are consistent with an evolutionary scenario in which the three types of dwarf galaxy have a close relation with each other. ", "introduction": "Although the faint end of the galaxy luminosity function is not well determined, recent studies indicate that low-surface-brightness (LSB) dwarf galaxies are by far the most numerous type of galaxy, and contribute a significant fraction of the mass of the universe (Reaves 1983; Binggeli, Sandage, \\& Tammann 1985; Phillipps et al.\\ 1987)\\markcite{reaves1983, sandage1985, phillipps1987}. Morphologically, dwarf galaxies, like their counterpart bright galaxies, are classified into several types. The most common type of dwarf galaxy ($\\sim 80\\%$ of the total) is the dwarf elliptical (dE). These galaxies have regular elliptical isophotes and roughly exponential surface brightness profiles; they are often found in groups and clusters (Davies et al.\\ 1988).\\markcite{davies1988} The second type of dwarf galaxy is the blue compact dwarf (BCD) galaxy. In contrast to gas-poor dEs, BCDs contain giant HII regions surrounding O and B stars within a massive HI reservoir; BCDs exhibit spectra slowly rising toward the blue, implying that they are undergoing intense star formation (du Puy 1970; Searle \\& Sargent 1972).\\markcite{dupuy1970, searle1972} Most BCDs have regular isophotes in the outer region, like dEs, but the inner isophotes are frequently distorted from ellipses, due to the presence of bright HII regions (Loose \\& Thuan 1986).\\markcite{loose1986} The final type of dwarf galaxy is the LSB dwarf galaxy. LSB dwarfs include both irregular (dI) and more regular spiral (dS) galaxies. Like BCDs, they contain a large amount of HI, often with small OB associations, and have blue colors ($B-V\\sim 0.5\\ {\\rm mag}$), indicating a significant level of recent star formation (Staveley-Smith, Davies, \\& Kinman 1992)\\markcite{staveley1992}. However, they are distinguished from BCD galaxies by their amorphous shapes even in the outer region. Additionally, in contrast to dEs, these galaxies are more likely to be found outside of clusters (Bingelli, Tarenghi, \\& Sandage 1990).\\markcite{bingelli1990} The evolutionary connections among the three different types of dwarf galaxies remain both elusive and confusing. There are two major competing hypotheses for the evolutionary connection between BCDs and dEs. The first hypothesis claims that BCDs are basically a different population from dEs, as evidenced by the spectroscopic and spectrophotometric differences. According to this scenario, BCDs are truly young systems, in which the present star burst is the first in the galaxy's lifetime. The second hypothesis suggests that BCDs, like dEs, are mainly composed of old stellar populations, and that their observed spectroscopic features and spectral energy distributions are the result of a recent burst of star formation (Staveley-Smith et al.\\ 1992).\\markcite{stavely1992} As an evidence for the second scenario, it is argued that the near-infrared emission in the vast majority of BCDs is attributable to old K and M giants, which are the major component of dEs (Thuan 1983; Hunter \\& Gallagher 1985).\\markcite{thuan1983, hunter1985} Similarly, there exist two competing hypotheses to explain the evolutionary connection between dIs and dEs. The first hypothesis states that dE galaxies are the faded remnants of previously actively star-forming dI galaxies whose gas has been lost. There exists circumstantial evidence that dEs have in fact evolved directly from dIs. Faber \\& Lin (1983)\\markcite{faber1983} and Kormendy (1985)\\markcite{kormendy1985} have used the similarity in the surface brightness profiles of dIs and dEs, which are mostly exponential, to argue that gas-rich dIs are the progenitors of dEs. The second hypothesis for the relation between dIs and dEs states that they represent parallel sequences of dwarf galaxies, fundamentally separated by the intrinsic difference in their structure. The observational evidence for this hypothesis is based mostly on the differences in appearance between the two types of dwarf galaxies; for instance, dIs have a more diffuse light distribution than dEs, and lack the bright nucleation which is frequently found in dEs. In addition to these differences, there is a dissimilarity in the flattening distribution of dEs and dIs; the apparent flattening of a galaxy is customarily given either by the apparent axis ratio $q$ or by the ellipticity $\\epsilon \\equiv 1 - q$. Bothun et al.\\ (1986) and Impey \\& Bothun (1997)\\markcite{bothun1986, impey1997} presented the results of Ichkawa, Wakamatsu, \\& Okamura (1986)\\markcite{ ichikawa1986} and Caldwell (1983)\\markcite{caldwell1983} as evidence for the different flattening distributions between dEs and dIs. However, the analysis of Ichikawa et al.\\ was based on the comparison between the flattening distributions of dEs and bright (non-dwarf) spiral galaxies; the situation is similar for Caldwell's analysis. In addition, contrary to the claims of Bothun et al.\\ and Impey \\& Bothun, both Ichikawa et al.\\ and Caldwell showed that the flattening distribution of dEs is similar to that of bright irregular galaxies. There have been previous attempts to compare the apparent axis ratio distributions between LSB dI galaxies and other types of dwarf galaxies. For example, Staveley-Smith et al.\\ (1992)\\markcite{stavely1992} constructed the axis ratio distribution for 438 Uppsala Galaxy Catalogue (hereafter UGC, Nilson 1973)\\markcite{nilson1973} LSB galaxies, and compared it to that of BCDs whose ellipticities were measured from Palomar Observatory Sky Survey (POSS) plates by Gorden \\& Gottesman (1981).\\markcite{gorden1981} However, previous studies of axis ratio distributions suffer from large uncertainties for several reasons. First, owing to the small dimensions and low surface brightness of dwarf galaxies, estimating their axis ratio is difficult and leads to large uncertainties. Second, the UGC sample used by Staveley-Smith et al.\\ (1992)\\markcite{stavely1992} is known to be inhomogeneous, containing galaxies ranging from true dwarf galaxies to more luminous very low surface brightness systems (Thuan \\& Seitzer 1979; McGaugh, Schombert, \\& Bothun 1995).\\markcite{thuan1979, mcgaugh1995} Finally, previous determinations of LSB dI axis ratios have been based on photographic plates; for comparison with recent CCD observations of other types of dwarf galaxy, it is essential to have measurements of the axis ratios of a homogeneous sample of LSB dIs based on modern CCD observations. In this paper, we measure the ellipticities of 30 LSB dI galaxies and compare the ellipticity distribution with that of 80 dEs (Ryden \\& Terndrup 1994; Ryden et al.\\ 1998)\\markcite{ryden 1994; 1998} and 62 BCDs (Sung et al.\\ 1998).\\markcite{sung1998} We find that the ellipticity distribution of LSB dIs is very similar to that of BCDs, and marginally different from that of dEs. We then determine, under various assumptions, the distribution of intrinsic shapes of dI galaxies and compare it to that of other types of dwarfs. First, we assume that LSB dIs are either all oblate or all prolate, and use non-parametric analysis to find the best-fitting distribution of intrinsic shapes. With this assumption, we find that the scarcity of nearly circular LSB dIs implies, at the 99\\% confidence level, that they cannot be a population of randomly oriented oblate or prolate objects. Next, we assume that dIs are triaxial, and use parametric analysis to find permissible distributions of intrinsic shapes. We find that if the intrinsic axis ratios, $\\beta$ and $\\gamma$, are distributed according to a Gaussian with means $\\beta_0$ and $\\gamma_0$ and a common standard deviation of $\\sigma$, the best-fitting set of parameters for LSB dIs is $(\\beta_0,\\gamma_0,\\sigma) = (0.66,0.50,0.15)$, and the best fit for BCDs is $(\\beta_0,\\gamma_0,\\sigma) = (0.66,0.55,0.16)$, while the best fit for dEs is $(\\beta_0,\\gamma_0,\\sigma) = (0.78,0.69,0.24)$. The LSB dIs and BCDs thus have a very similar shape distribution, given this triaxial hypothesis, while the dEs peak at a somewhat more spherical shape. Our results are consistent with an evolutionary scenario in which the three types of dwarf galaxy have a close relation with each other. ", "conclusions": "We measure the ellipticities for a sample of 30 LSB dIs and compare the distribution of ellipticities with those for the samples of 62 BCDs and 80 dEs. From this comparison, we find that the axis ratio distribution of LSB dIs is very similar to that of BCDs. Compared to dEs, LSB dIs are slightly flatter, but the difference is marginal. We also determine the intrinsic shape of LSB dIs from the distribution of apparent axis ratios. From the non-parametric analysis, we find the hypothesis that our sample LSB dIs are randomly oriented oblate or prolate objects is rejected with strong confidence level. On the other hand, the shape of LBS dI galaxies are well described by triaxial spheroids if their axis ratios, $\\beta$ and $\\gamma$, have a Gaussian distribution. From the parametric analysis, we determine the best-fitting parameters are $(\\beta_0,\\gamma_0,\\sigma) = (0.66, 0.50, 0.15)$. These results directly contradict the long-standing belief that LSB dIs have very flattened disky shapes, quite different from the spheroidal shapes of dEs and BCDs. Therefore, our results are consistent with the scenario that the three major types of dwarf galaxies have very close evolutionary connections." }, "9803/astro-ph9803025_arXiv.txt": { "abstract": " ", "introduction": "The 48~ms radio pulsar PSR~B1259$-$63 is in a highly eccentric 3.4~yr orbit with a Be star (Johnston et al. 1992). Near periastron, for a sufficiently strong Be star wind, PSR~B1259$-$63 could possibly make a transition to an accretion regime, exhibiting X-ray pulsations of luminosity $>$10$^{35}$~erg~s$^{-1}$. {\\it ASCA} X-ray observations at three epochs around the 1994 periastron passage (Kaspi et al. 1994; Hirayama 1996) found the source to be unpulsed, with modest luminosity $\\sim$10$^{34}$~erg~s$^{-1}$, for a distance of 2~kpc. These properties imply that shock emission produces the observed X-rays, rather than accretion (Tavani \\& Arons 1997). However radio timing observations suggested sudden, brief accretion events near periastron that could not be ruled out by the X-ray observations (Manchester et al. 1995). Indeed the pulsar's anomalously low period suggests it may have undergone occasional accretion episodes in the past. {\\it ASCA} observations during the 1997 periastron passage were impossible due to solar constraints. The ASM on {\\it RXTE} has been observing bright celestial X-ray sources regularly since early 1996. The instrument consists of three ``scanning shadow cameras,'' each containing a position-sensitive proportional counter that is mounted below a wide-field collimator, covered by a coded mask. The instrument provides roughly 5 celestial scans per day, with diminished exposure in directions toward the Sun. Further information on the ASM is given by Levine et al. (1996). While PSR~B1259$-$63 is not routinely monitored for the ASM source-history database, we extracted an X-ray light curve as an archival analysis project using standard techniques. The derived light curve (Figure~1a) covers the time interval of MJD 50178--50762 (1996 Apr 5 to 1997 Nov 10). The overall mean intensity of PSR B1259$-$63 during this time interval was $(0.5 \\pm 2.4)\\times 10^{-11}$~erg~s$^{-1}$~cm$^{-2}$ at 2--10 keV. This result is given with consideration of both the systematic bias and uncertainty in the ASM light curves for faint, yet fairly isolated X-ray sources. Figure~1b shows $3\\sigma$ upper limits obtained in the periastron vicinity in 2-day averages. It shows that in the 2 weeks following periastron, the source was not observed to be brighter than $\\sim$10 times the 1994 periastron luminosity. The $3\\sigma$ upper limits to the X-ray emission from PSR~B1259$-$63 from the ASM thus argue that the pulsar did not undergo even brief episodes of accretion during the 1997 periastron passage. This is consistent with the shock emission model, as well as with new conclusions based on timing that rule out the sudden ``spin-ups'' that were claimed previously (Wex et al. 1998). \\begin{figure}[t] \\centerline{ \\psfig{figure=psr1259.ps,height=5.5cm} \\psfig{figure=psr1259_upper.ps,height=5.5cm} } \\caption{(a) ASM light curve for PSR~B1259$-$63. (b) 3$\\sigma$ upper limits to the ASM flux near periastron from (a). The horizontal dashed line represents the brightest {\\it ASCA} 1994 periastron detection (Kaspi et al. 1994), and the vertical dotted line indicates periastron.} \\end{figure} ", "conclusions": "The 48~ms radio pulsar PSR~B1259$-$63 is in a highly eccentric 3.4~yr orbit with a Be star (Johnston et al. 1992). Near periastron, for a sufficiently strong Be star wind, PSR~B1259$-$63 could possibly make a transition to an accretion regime, exhibiting X-ray pulsations of luminosity $>$10$^{35}$~erg~s$^{-1}$. {\\it ASCA} X-ray observations at three epochs around the 1994 periastron passage (Kaspi et al. 1994; Hirayama 1996) found the source to be unpulsed, with modest luminosity $\\sim$10$^{34}$~erg~s$^{-1}$, for a distance of 2~kpc. These properties imply that shock emission produces the observed X-rays, rather than accretion (Tavani \\& Arons 1997). However radio timing observations suggested sudden, brief accretion events near periastron that could not be ruled out by the X-ray observations (Manchester et al. 1995). Indeed the pulsar's anomalously low period suggests it may have undergone occasional accretion episodes in the past. {\\it ASCA} observations during the 1997 periastron passage were impossible due to solar constraints. The ASM on {\\it RXTE} has been observing bright celestial X-ray sources regularly since early 1996. The instrument consists of three ``scanning shadow cameras,'' each containing a position-sensitive proportional counter that is mounted below a wide-field collimator, covered by a coded mask. The instrument provides roughly 5 celestial scans per day, with diminished exposure in directions toward the Sun. Further information on the ASM is given by Levine et al. (1996). While PSR~B1259$-$63 is not routinely monitored for the ASM source-history database, we extracted an X-ray light curve as an archival analysis project using standard techniques. The derived light curve (Figure~1a) covers the time interval of MJD 50178--50762 (1996 Apr 5 to 1997 Nov 10). The overall mean intensity of PSR B1259$-$63 during this time interval was $(0.5 \\pm 2.4)\\times 10^{-11}$~erg~s$^{-1}$~cm$^{-2}$ at 2--10 keV. This result is given with consideration of both the systematic bias and uncertainty in the ASM light curves for faint, yet fairly isolated X-ray sources. Figure~1b shows $3\\sigma$ upper limits obtained in the periastron vicinity in 2-day averages. It shows that in the 2 weeks following periastron, the source was not observed to be brighter than $\\sim$10 times the 1994 periastron luminosity. The $3\\sigma$ upper limits to the X-ray emission from PSR~B1259$-$63 from the ASM thus argue that the pulsar did not undergo even brief episodes of accretion during the 1997 periastron passage. This is consistent with the shock emission model, as well as with new conclusions based on timing that rule out the sudden ``spin-ups'' that were claimed previously (Wex et al. 1998). \\begin{figure}[t] \\centerline{ \\psfig{figure=psr1259.ps,height=5.5cm} \\psfig{figure=psr1259_upper.ps,height=5.5cm} } \\caption{(a) ASM light curve for PSR~B1259$-$63. (b) 3$\\sigma$ upper limits to the ASM flux near periastron from (a). The horizontal dashed line represents the brightest {\\it ASCA} 1994 periastron detection (Kaspi et al. 1994), and the vertical dotted line indicates periastron.} \\end{figure}" }, "9803/astro-ph9803163_arXiv.txt": { "abstract": "We describe a concept for an imaging spectrograph for a large orbiting observatory such as NASA's proposed Next Generation Space Telescope (NGST) based on an imaging Fourier transform spectrograph (IFTS). An IFTS has several important advantages which make it an ideal instrument to pursue the scientific objectives of NGST. We review the operation of an IFTS and make a quantitative evaluation of the signal-to-noise performance of such an instrument in the context of NGST. We consider the relationship between pixel size, spectral resolution, and diameter of the beamsplitter for imaging and non-imaging Fourier transform spectrographs and give the condition required to maintain spectral modulation efficiency over the entire field of view. We give examples of scientific programs that could be performed with this facility. ", "introduction": "The Association of Universities for Research in Astronomy (AURA), with NASA support, recently appointed a committee to ``study possible missions and programs for UV-Optical-IR astronomy in space for the first decades of the twenty-first century.'' The report urged the development of a general-purpose, near-infrared observatory equipped with a passively cooled primary mirror ($T \\le 70$~K) with a minimum diameter of 4 meters (Dressler 1996). To enhance its performance, the report recommended that the observatory be placed as far from the Earth-Moon system as possible to reduce stray light and to maintain the telescope's relatively low temperature. With such a facility, it should be possible to learn in detail how galaxies formed, measure the large-scale curvature of space-time by measuring distant standard-candles, trace the chemical evolution of galaxies, and study nearby stars and star-forming regions for signs of planetary systems. A detailed discussion of the next generation space telescope (NGST) and its scientific potential given by Stockman (1997). For NGST to attain these scientific objectives, it must have an instrument which is designed to execute panchromatic observations over the critical 1--15 $\\mu$m wavelength range of the faintest detectable objects. With nJy sensitivity levels attainable at near-infrared (1--5 $\\mu$m) (NIR) and mid-infrared (5--15$\\mu$m) (MIR) wavelengths, NGST will be able to study the well-calibrated rest-frame optical diagnostics in distant ($z=3-10$) galaxies, thus probing for the first time, their stellar content, star-formation history and nuclear activity. At the longer wavelengths, NGST can investigate these properties in $z=3-5$ galaxies using diagnostics that are unaffected by dust extinction and reddening, and also study the dust properties directly. At the flux limits characteristic of NGST, the confusion limit is likely to be approached, with virtually every pixel having significant information (e.g., by extrapolation from counts in the Hubble Deep Field (Williams et al. 1996)). As a result, one of the best ways to maximize the scientific output from NGST is to provide a wide-field imaging spectrograph that is efficient in this limit. An imaging Fourier transform spectrometer (IFTS) provides these capabilities in a low-cost, high throughput, compact design. It provides the only efficient means of conducting {\\it unbiased} spectroscopic surveys of the high-$z$ Universe, i.e., without object preselection (e.g., using broad band colors) and without the restrictions imposed by spectrometer slit geometry and placement. An IFTS also allows spectroscopy over a wide bandpass, affords flexibility in choice of resolution, is easy to calibrate, and is ideal for wide-field spectroscopic surveys. Bennett et al. (1993) and Bennett, Carter, \\& Fields (1995) describe the operating principles of imaging Fourier transform spectrographs and compare their performance with alternative imaging spectrometers. A comprehensive review of the application of interferometers and the techniques of Fourier spectroscopy to astrophysical problems is given by Ridgway \\& Brault (1984), and a recent summary of the field, including a description of an astronomical IFTS is given by Maillard (1995). Spaceborne Fourier transform spectrometers have been responsible for spectacular results in the fields of planetary exploration and cosmology. Infrared FT spectrometers developed at the Goddard Space Flight Center (GSFC) flew on board the Mariner 9 mission to Mars, and were carried to the outer planets by the Voyager spacecraft (Hanel et al. 1992). The instruments provided superb data revealing, for the first time, the composition of the atmospheres of the giant gaseous planets (e.g. Jupiter, Hanel et al. 1979). The Composite Infrared Spectrometer (CIRS), currently traveling to Saturn onboard the Cassini spacecraft, is another instrument developed at GSFC. CIRS is the first step towards an imaging FTS as it has a linear array of detectors, rather than a single element detector, in order to map the temperature and composition of the atmospheres of Saturn and Titan as a function of altitude during limb soundings. (Kunde et al. 1996). The definitive measurement of the spectrum of the cosmic microwave background radiation (CMBR) was one of the most dramatic experimental measurements of this decade (Mather et al. 1990; Gush, Halpern \\& Wishnow 1990). The FIRAS instrument onboard the NASA satellite COBE which first performed this measurement, and the COBRA rocket experiment conducted by the University of British Columbia which confirmed it a few months later, were both liquid-helium cooled, differential Fourier transform spectrometers. These instruments used a dual-input, dual-output configuration where one input viewed the sky and the other viewed a blackbody calibrator (Mather, Fixen \\& Shafer 1993; Gush \\& Halpern 1992). Absolute photometric measurements were obtained by reference to the blackbody calibrator, and the CMBR was observed to have an undistorted Planck spectrum corresponding to a temperature of $2.728\\pm0.004$ K (Fixen et al. 1996). The IFTS proposed here can be thought of as an extension of these experiments where focal plane detector arrays yield simultaneous imaging and spectral information. In the next decade, missions such as WIRE, AXAF, and SIRTF will expand astrophysical horizons, possibly unveiling entirely new populations of objects. An IFTS offers the flexibility (e.g., spectral resolution) that may prove essential in investigating the nature of these sources. Due to its flexibility and its ability to provide simultaneous imaging and spectroscopy of every object in the field of view (FOV), an IFTS is a {\\it necessary} instrument for the NGST mission. \\section {IFTS CONCEPT} \\label{concept} An IFTS (Fig. \\ref{fts}) is axis-symmetric, and the optical path difference (OPD) is the same for all the points of the image with the same angle of incidence from the axis of the interferometer. Hence, the FOV is circular. On the object side, an entrance collimator illuminates the interferometer with parallel light. The interfering beams are collected by the output camera, creating a stigmatic relation between the object and image planes. By placing a detector array in the output focal plane the entrance field is imaged on the array and each pixel works as a single detector matched to a point on the sky. \\begin{figure}[thb] \\centerline{\\psfig{figure=fig1.eps,width=3.3in,angle=-90}} \\caption{\\small A sketch of the optics of a simple single-beam imaging Fourier transform spectrograph consisting of a collimating lens, a beam splitter, two mirrors (one movable), and a camera lens. The optical path difference is $x$.} \\label{fts} \\end{figure} Retrieving spectral information involves recording the interferogram generated by the source imaged onto the focal plane array (FPA). The OPD is scanned in discrete steps since FPAs are integrating detectors. Scanning in this way generates a data cube of two-dimensional interferograms. The signal from the same pixel in each frame forms an independent interferogram. These interferograms are Fourier transformed individually yielding a spectral data cube composed of the same spatial elements as the image. \\subsection{A Perfect Match to NGST Science} \\label{perfect} The features of an IFTS which make it the instrument of choice for NGST are efficiency, flexibility, and compactness. The most compelling reason for choosing an IFTS is that in the dual port design (see Fig. \\ref{optical-layout}) virtually every photon collected by the telescope is directed towards the focal plane for detection. Other solutions are inefficient, inflexible, and wasteful of mass, power, and volume. Cameras equipped with filters admit only a restricted bandpass at low spectral resolution. To compete with the spectral multiplex advantage of an IFTS, a camera system needs multiple dichroics and FPAs. The additional mass and thermal load is a severe penalty. Classical dispersive spectrographs have slit losses, grating inefficiencies due to light lost in unwanted orders, and limited free spectral range (the same is true for a Fabry-Perot). {\\em An IFTS acquires full bandpass imaging simultaneously with higher spectral resolution data.} Therefore, a high SNR broad-band image always accompanies full spectral sampling of the FOV with no penalty in integration time. An IFTS is a true imaging spectrograph and measures a spectrum for every pixel in the FOV. It is not necessary to choose which regions in the image are most deserving of spectroscopic analysis. Overheads are eliminated because no additional observing time is needed for imaging prior to object selection, and there is no delay in positioning slit masks, fibers, or image slicing micro-mirrors. Thus, an IFTS will produce a rich scientific legacy with tremendous potential for serendipity. \\begin{figure}[thb] \\centerline{ \\psfig{figure=fig2.ps,width=3.3in,angle=-90}} \\caption{\\small Schematic optical layout of a 60$^\\circ$ dual-input, dual-output Michelson interferometer. } \\label{optical-layout} \\end{figure} Table \\ref{capabilities} details the capabilities of a putative IFTS suitable for NGST. We use the instrument described by this table to illustrate the potential of an IFTS. Two points in Table \\ref{capabilities} must be stressed: 1) An IFTS is spectrally multiplexed, therefore all spectral channels are obtained simultaneously within the stated integration time. 2) The free spectral range of an IFTS is limited only by the band-pass filter and the detector response. Consequently, the usual definition of resolution, $R = \\lambda/\\delta \\lambda$, is of limited use. It is conventional to scan the OPD of an IFTS in equal steps so that the resolution is constant in wavenumber, $k$. Thus, we use $M$ to denote the number of spectral channels. For example, in the NIR with a 1-5 $\\mu$m band-pass, $M=5$ means that $\\delta k = (k_{max} - k_{min})/M = 1600$~cm$^{-1}$, and a scan yields 5 bands centered 1.1, 1.3, 1.7, 2.3, \\& 3.6 $\\mu$m. \\begin{deluxetable}{lll} \\tablewidth{0pt} \\tablecaption{Capabilities of a Putative NGST IFTS} \\tablehead{ \\colhead{} & \\colhead{NIR Channel} &\\colhead{MIR Channel}} \\startdata Design\t\t\t& Dual-port & Dual-port \\\\ Bandpass\t\t& 1-5 $\\mu$m & 5-15 $\\mu$m \\\\ Resolution \t\t& 1 cm$^{-1}$\t& 1 cm$^{-1}$ \\\\ FOV \t\t\t& $200''$\t& $100''$ \\\\ Pixel size \t\t& $0.''05$\t& $0.''1$ \\\\ Array format \t\t& 4k$\\times$4k & 1k$\\times$1k \\\\ Detector\t\t& InSb \t\t& HgCdTe \\\\ Throughput\t\t& $> 0.5$\t& $> 0.5$ \\\\ Sensitivity\\tablenotemark{a} \\\\ ~$M=1$\t\t& 200 pJy & 13 nJy \\\\ ~$M=5$\t\t\t& 1 nJy & 65 nJy \\\\ ~$M=100$\t\t& 35 nJy & 1.3 $\\mu$Jy \\\\ \\enddata \\tablenotetext{a}{SNR = 10 for a $10^5$~s integration over the entire spectral band for a point source. $M$ is the number of simultaneous spectral channels in the band-pass --- see \\S \\ref{perfect}. Note that all spectral channels are obtained simultaneously. The spectrum is assumed to be flat in $F_\\nu$ and the SNR is quoted at 2 $\\mu$m for the NIR channel, and at 10 $\\mu$m for the MIR channel.} \\label{capabilities} \\end{deluxetable} The throughput of an IFTS with ideal optics is only limited by the efficiency of the beam splitter. In a dual-input, dual-output port design no light is wasted and the throughput approaches 100\\%. An IFTS has no loss of light or spatial information because there is no slit, hence an IFTS is perfectly adapted to doing multi-object spectroscopy in crowded or confusion limited fields. A IFTS uses every photon whereas traditional cameras and spectrographs throw away photons (either spectrally with a filter or spatially with a slit), so at a very fundamental level an IFTS is superior. On blaze, a good grating is 80\\% efficient, but averaged over the free spectral range this drops to about 65\\%. An IFTS is not optimized for single-object spectroscopy because the broad-band photon shot noise is associated with every frame in the interferogram. Hence, for a single object a slit spectrograph is $\\eta_g \\eta_s M$ times faster than an IFTS of the same resolution in background limited operation, where $\\eta_g$ is the grating efficiency averaged over the blaze function, and $\\eta_s$ is the slit loss, where typically the product $\\eta_g \\eta_s \\approx 0.3$. This disadvantage is more than compensated for by the spatial-multiplexing capability of an IFTS. A typical deep background limited exposure of an IFTS will reach $K=29.5$, SNR=10 and will contain at least 3500 and possibly, depending on cosmology, up to 11,000, objects per field (see Fig. \\ref{number-counts}). A grating spectrograph with a fiber feed or multi-slit capability can perhaps record spectra for only a few percent of these objects at a time, requiring hundreds of pointings to make an unbiased survey of a single field, as opposed to the single IFTS imaging-cum-spectroscopic observation. \\begin{figure}[thb] \\centerline{\\psfig{figure=fig3.ps,width=3.4in,angle=90}} \\caption{\\small $K$-band number counts (Djorgovski et al. 1995; Gardner et al. 1993, 1996; Glazebrook et al. 1994; Huang et al. 1997; Mobasher et al. 1986; Moustakis et al. 1997; McLeod et al. 1995; \\& Metcalfe et al. 1996) together with models of the luminosity function modeled using the formalism of Gardner (1998), which has been used to extrapolate the number counts into the NGST domain. The solid lines include the effects of passive evolution, while the dashed lines include only K-corrections. The upper line in each case is for $q_0 = 0.1$, and the lower lines are for $q_0 = 0.5$. Current number counts imply at least 3500 objects per $3.'3$ NGST field, while the extrapolations shown here suggest as many as 11,000 to $K=29.5$. } \\label{number-counts} \\end{figure} An IFTS is tolerant of detector noise because it always operates under photon limited conditions due to the broad spectral bandpass transmitted to the FPA. This is illustrated in Table \\ref{readoutexamples} which shows a break-down of the noise sources in the NIR and MIR channels corresponding to the performance listed in Table \\ref{capabilities}. Table \\ref{readoutexamples} also shows that the read-out rates required to avoid saturation are modest ($1-10$~mHz), since typical well depths for NIR InSb or HgCdTe arrays are a few $10^5$ e$^-$ and $10^7$ e$^-$ for MIR Si:As arrays. \\begin{deluxetable}{llll} \\tablewidth{0pt} \\tablecaption{Signal and Noise Budget\\tablenotemark{a}} \\tablehead{ \\colhead{} & \\colhead{} & \\colhead{NIR Channel} &\\colhead{MIR Channel}} \\startdata & & $F$ = 1 nJy & $F$ = 65 nJy \\\\ & & $t$ = 1000 s & $t$ = 100 s \\\\ & & $\\Delta\\lambda = 1-5\\mu$m & $\\Delta\\lambda = 5-15\\mu$m \\\\ \\hline Signal \\\\ & Source & 610 & 2709 \\\\ & Background\\tablenotemark{b} & 3724 & 463669 \\\\ & & & \\\\ Total Signal & & 4334 & 466378 \\\\ \\\\ Noise\\\\ & Signal Shot Noise & 24.7 & 52.0 \\\\ & Background Shot Noise & 60.8 & 680.9 \\\\ & Dark Shot Noise & 5.5 & 10.0 \\\\ & Read Noise & 5 & 5 \\\\ Total Noise & & 66.0 & 683.0 \\\\ \\enddata \\label{readoutexamples} \\tablenotetext{a}{In electrons} \\tablenotetext{b} {Background includes zodiacal foreground and thermal emission from the telescope as calculated as described in \\S \\ref{snrcal}.} \\end{deluxetable} Similarly, orders of magnitude higher thermal emission from the instrument, or thermal radiation leaks from outside the instrument bay, can be tolerated compared to the case for dispersive spectrometers or fixed filter cameras. As a pragmatic demonstration of this principle, the IFTS instruments LIFTIRS and HIRIS are routinely operated with ambient temperature optics in the 8-14 $\\mu$m band (Bennett et al. 1995), whereas dispersive spectrometers, like SEBASS (Bennett, {\\it private communication}), operating in the same spectral region must have the slit and all following optics cooled far below ambient temperatures. The reason is that in a dispersive spectrometer the thermal emission of all the elements and optics downstream of the slit reach the detector at full spectral range determined by the bandpass limiting element at or near the coldstop, whereas only the narrow spectral range corresponding to the width of a spectral channel for the signals of interest reach the detector pixels. For the IFTS, both the signals of interest and the thermal emission are seen over the full spectral range determined by the bandpass limiting filter, and thus it is only necessary that the thermal emission of the optical elements along the optic axis integrated over the bandpass of interest, either 1-5 $\\mu$m or 5-15 $\\mu$m, be somewhat less than that of the integral of the zodiacal foreground, telescope emission, and source signal level integrated over the same broad spectral region. An IFTS is potentially immune to cosmic ray hits because the ``energy'' of a single upset pixel in one OPD frame appears as a sinusoidal signal divided among all bins in the spectral transform of the interferogram for that pixel. We can ignore cosmic ray hits only if the counts generated are at or below our noise level. A minimum ionizing cosmic ray proton ($E \\simeq 1$ GeV) has ionization losses of $dE/dx \\simeq 400$ eV $\\mu$m$^{-1}$ in Si. Assuming that 3.6 eV is required to produce an electron-hole pair, a cosmic ray will yield at least a few thousand events, since typical pixels have sensitive layers that are tens of $\\mu$m thick. We would obtain similar number for a hybrid device, i.e., InSb or HgCdTe on a Si multiplexer. If a cosmic ray hit produces a significant signal in a certain number of pixels, those pixels must be ``repaired'' by interpolating the interferogram between the previous few ``good'' frames, and the following few ``good'' frames which are not contaminated by cosmic ray hits. The same sort of processing would be needed for any other system as well, be it an imager or a spectrometer. Comparison with the noise sources listed in Table \\ref{readoutexamples} indicates that cosmic ray hits will have to be repaired in the NIR channel, while the MIR channel will be more tolerant. A dual port design (Fig. \\ref{optical-layout}) delivers the complementary symmetric and antisymmetric interferograms. In this dual-input dual-output design, the field of the complementary input (labeled ``Calibration Input'' in Fig. \\ref{optical-layout}) is also imaged and superimposed on each image of the ``Primary Input''. This property is often used to cancel the sky emission. In operation, when observing the sky in the primary input, the secondary input would be fed with a cold blackbody load, having negligible radiance. The final interferogram is constructed from the difference between the two outputs (which is therefore also immune to common mode electrical noise) while the normalized ratio reveals systematic variation due to detector drifts. The wavelength scale and the instrumental line shape (a $sinc$ function if there is no apodizing) are precisely determined and are independent of wavelength. Absolute wavelength calibration is done by counting fringes of an optical single-mode laser. Compared to a dispersive system the broad-band operation of an IFTS means that there are $M$ times more photons for flat fielding and determining signal-dependent gain (linearity). Hence, high signal-to-noise calibration images can be acquired faster or with lower power internal sources. \\subsection{Pixel Size, Spectral Resolution, and Field of View} \\label{pixelsize} Spatial multiplexing renders the performance of an IFTS equal to that of an ideal multi-slit spectrograph (Bennett 1995). Hence, even if we ignore slit losses and blaze inefficiency the other advantages of an IFTS are overwhelming. The spectral resolution can be varied arbitrarily from the coarsest case of a small number of bands up to a spectral resolution limit determined only by the maximum OPD characteristic of the instrument. The proposed instrument has a maximum OPD of 1 cm and hence can operate over a range of resolutions from full band up to $M$=8000 in the NIR The spectral resolution limit, $R = k/\\delta k$, of a Michelson interferometer is \\begin{equation} R = 8(d/ \\phi D)^2, \\end{equation} \\noindent where $\\phi$ is the angular diameter of the FOV, $d$ is the diameter of the beam splitter, and $D$ the telescope primary mirror diameter (e.g., Jacquinot 1954; Maillard 1995). Classically, $\\phi$ refers to the entire field, but in the case of an IFTS, $\\phi$ is the FOV of an on-axis pixel. Although convenient if a single fringe fills the FPA, just as with imaging Fabry-Perots, there is no reason why each pixel should record the same apparent wavenumber. Fringes crowd together with increasing field angle. Therefore, the need to maintain modulation efficiency over the entire field of view requires that the spatial separation of the fringes at the edge of the FPA, for a given retardance, is significantly greater than the pixel spacing. If $x$ is the OPD for a normally incident beam with wavenumber $k$, and $\\theta$ is the field angle of off-axis rays at the beam splitter, then the path difference at $\\theta$ is $x_\\theta = x cos\\theta$ and the apparent wavenumber of this beam is \\begin{equation} k_\\theta = k/cos\\theta. \\end{equation} \\noindent The angles $\\theta$ and $\\phi$ are related by the angular magnification, $D/d$. If $\\delta \\theta$ is the angular width, also at the beam splitter, corresponding to a single pixel, the spectral resolution limit for off-axis points can be found by differentiating Eq (2), \\begin{equation} 1/R_\\theta = \\delta k_\\theta/k_\\theta = tan\\theta \\delta \\theta, \\end{equation} \\noindent Fig. \\ref{pixel_fov} shows the pixel size for a given field of view for a range of resolutions. For example, for an 8~m diameter primary aperture and a beam splitter of diameter 10~cm, a FOV of $200''$, and a pixel size of $0.''05$, the corresponding angles at a beam splitter of diameter 10~cm are $2.^\\circ 2$ and $4''$ respectively, leading to a resolution limit of $R = 1.3 \\times 10^6$. Since this resolution is two orders of magnitude greater than we are proposing, it is clear that spectral resolution is not the principal factor determining pixel size. An alternative way to view this constraint is that $d$, i.e., the size of the optics, is determined not by spectral resolution, but by the requirement that there be no vignetting over the field of view. Thus, the optics for an IFTS are similar to that of a simple re-imaging camera, and are smaller and slower than those of an equivalent dispersive spectrograph. \\begin{figure}[thb] \\centerline{\\psfig{figure=fig4.ps,width=3.4in,angle=90}} \\caption{\\small Pixel size as a function of the field of view required to spatially fully sample fringes at the edge of the FPA, and hence maintain modulation efficiency. Curves are plotted for resolutions $R = k/\\delta k = 10^4$, $10^5$ and $10^6$ assuming an 8~m diameter primary aperture and a beam splitter of diameter 10~cm. } \\label{pixel_fov} \\end{figure} We therefore have broad freedom to choose the pixel size by trading off field of view and spatial sampling. Given that NIR arrays of $4096 \\times 4096 $ pixels are likely to be available in the near future, a pixel size of $0.''05$ yields a $3.'3$ field of view and $\\lambda / 2D$ sampling at 4 $\\mu$m. This choice of pixel size does not preclude diffraction limited imaging at shorter wavelengths. If pixels have sharp boundaries, then it is possible to extract information at spatial frequencies above the cut-off in the pixel-sampling modulation transfer function if the spacecraft can offset and track at the sub-pixel level (cf. Fruchter and Hook 1998). Similar reasoning suggests $0.''1$ pixels would be a satisfactory compromise for the MIR channel. ", "conclusions": "An IFTS instrument can perform a wide variety of NGST science. The advantages of the IFTS concept are: \\begin{enumerate} \\item Deep imaging acquired simultaneously with higher spectral resolution data over a broad wavelength range. \\item ``Hands-off'', unbiased, multi-object, slitless spectroscopy (ideal for moving objects). Efficient in confusion limit. \\item Flexible resolution ($M=1-10,000$). \\item High throughput (near 100\\%) dual-port design. \\item Tolerant of cosmic rays, read-noise, dark current, and light leaks. \\item Simple and reliable calibration. High SNR determination of flat-fields and detector non-linearity. \\item Compact, lightweight design. Slow reimaging optics. \\end{enumerate}" }, "9803/astro-ph9803119_arXiv.txt": { "abstract": "We present a study of 581 Hz oscillations observed during a thermonuclear X-ray burst from the low mass X-ray binary (LMXB) 4U 1636-54 with the Rossi X-ray Timing Explorer (RXTE). This is the first X-ray burst to exhibit both millisecond oscillations during the rising phase as well as photospheric radius expansion. We measure an oscillation amplitude within 0.1 s of the onset of this burst of $75 \\pm 17 \\%$, that is, almost the entire thermal burst flux is modulated near onset. The spectral evolution during the rising phase of this burst suggests that the X-ray emitting area on the neutron star was increasing, similar to the behavior of bursts from 4U 1728-34 with 363 Hz oscillations reported recently. We argue that the combination of large pulsed amplitudes near burst onset and the spectral evidence for localized emission during the rise strongly supports rotational modulation as the mechanism for the oscillations. We discuss how theoretical interpretation of spin modulation amplitudes, pulse profiles and pulse phase spectroscopy can provide constraints on the masses and radii of neutron stars. We also discuss the implications of these findings for the beat frequency models of kHz X-ray variability in LMXB. ", "introduction": "Large amplitude millisecond oscillations have now been observed during thermonuclear X-ray bursts from six low mass X-ray binary (LMXB) systems with the Rossi X-ray Timing Explorer (RXTE) (see \\markcite{SZS}Strohmayer, Zhang \\& Swank 1997; \\markcite{SMB}Smith, Morgan \\& Bradt 1997, \\markcite{Z96}Zhang {\\it et al.} 1996; \\markcite{Swank97}Swank {\\it et al.} 1997; and \\markcite{Stroh97}Strohmayer {\\it et al.} 1997). The thermonuclear instability which triggers an X-ray burst burns in a few seconds the nuclear fuel which has been accumulated on the neutron star surface over several hours. This $>$ 10$^3$ difference between the accumulation and burning timescales means that it is extremely unlikely that the conditions required to trigger the instability will be achieved simultaneously over the entire stellar surface. This realization, first emphasized by \\markcite{Joss78}Joss (1978), led to the study of lateral propagation of the burning instability over the neutron star surface (see \\markcite{FW82}Fryxell \\& Woosley 1982, \\markcite{NIF84}Nozakura, Ikeuchi \\& Fujimoto 1984, and \\markcite{B95}Bildsten 1995). The subsecond risetimes of thermonuclear X-ray bursts suggests that convection plays an important role in the physics of the burning front propagation, especially in the low accretion rate regime which leads to large ignition columns (see \\markcite{B98}Bildsten (1998) for a review of thermonuclear burning on neutron stars). \\markcite{B95}Bildsten (1995) has shown that pure helium burning on neutron star surfaces is in general inhomogeneous, displaying a range of behavior which depends on the local accretion rate. Low accretion rates lead to convectively combustible accretion columns and standard type I bursts, while high accretion rates lead to slower, nonconvective propagation which may be manifested in hour long flares. These studies emphasize that the physics of thermonuclear burning is necessarily a multi-dimensional problem and that {\\it localized} burning is to be expected, especially at the onset of bursts. There is now good evidence that the oscillations seen during the rising phase of bursts from 4U 1728-34 are produced by spin modulation of such a localized thermonuclear hotspot on the surface of the neutron star, and that the observed oscillation frequency is a direct measure of the neutron star spin frequency (see \\markcite{SZS}Strohmayer, Zhang \\& Swank 1997). These observations provide the most compelling evidence to date that neutron stars in LMXB are rotating with near millisecond periods. In this Letter we present new burst data from the LMXB 4U 1636-54 which provides further evidence in support of the spin modulation hypothesis for the millisecond burst oscillations. We present data from a thermonuclear burst from 4U 1636-54 which reveals a strong transient oscillation during the burst rise at 1.723 ms with an initial amplitude of $75 \\pm 17 \\%$. We also discuss the implications of these findings for the spin modulation interpretation and how they can be used to place constraints on the mass and radius of neutron stars. Finally, we discuss some implications of our observations for the current theories of kilohertz quasiperiodic oscillations (QPO) in LMXB. ", "conclusions": "If the millisecond oscillations seen during bursts are in fact due to spin modulation, then detailed study and modelling of the oscillation amplitudes, pulse profiles and spectral variability with pulse phase during X-ray bursts can provide a wealth of information on the mass and radius of the neutron star. For example, the maximum modulation amplitude that can be obtained from a hotspot of a given angular size on a rotating neutron star is set by the strength of general relativistic light bending. For the case that rotation is not rapid enough to substantially distort the exterior spacetime from the Schwarzchild spacetime, and this is the case even for spin periods of a few milliseconds (Lamb \\& Miller 1995), then the maximum amplitude depends only on the compactness of the neutron star, that is, the ratio of stellar mass to radius $GM/c^2R$. Stars which are more compact produce lower amplitudes due to flattening of the pulse by light bending (see \\markcite{PFC}Pechenick, Ftaclas \\& Cohen 1983; \\markcite{S92}Strohmayer 1992; and \\markcite{ML97}Miller \\& Lamb 1997). Since the intrinsic rotational modulation amplitude can only be decreased by other effects such as photon scattering (see \\markcite{MLP}Miller, Lamb \\& Psaltis 1997; \\markcite{BL}Brainerd \\& Lamb 1987; and \\markcite{KP}Kylafis \\& Phinney 1989) or the viewing geometry of the spot, the maximum observed oscillation amplitude represents a lower limit to the intrinsic amplitude. Thus an observed amplitude can be used to place an upper limit on the compactness of the neutron star, that is, if the star were more compact than some limit it would not be able to produce a modulation amplitude as large as that observed. In principle, stellar rotation will also play a role in the observed properties of spin modulation pulsations. For example, assuming the oscillation frequency of 581 Hz represents the spin frequency of the neutron star in 4U 1636-54, then for a 10 km radius neutron star the spin velocity is $v_{spin}/c = 2\\pi \\nu_{spin} R \\approx 0.12$ at the rotational equator. The motion of the hotspot produces a Doppler shift of magnitude $\\Delta E / E \\approx v_{spin}/c = 0.12$, thus the observed spectrum is a function of pulse phase (see \\markcite{CS}Chen \\& Shaham 1989). Measurement of a pulse phase dependent Doppler shift in the X-ray spectrum would provide additional evidence supporting the spin modulation model and would also provide a means of constraining the neutron star radius. The rotationally induced velocity also produces an aberration which results in asymmetric pulses, thus the pulse shapes also contain information on the spin velocity and therefore the stellar radius (\\markcite{CS}Chen \\& Shaham 1989). The component of the spin velocity along the line of site is proportional to $\\cos\\theta$, where $\\theta$ is the lattitude of the hotspot measured with respect to the rotational equator. The modulation amplitude also depends on the lattitude of the hotspot, as spots near the rotational poles produce smaller amplitudes than those at the equator. Thus we expect a correlation between the observed oscillation amplitude and the size of any pulse phase dependent Doppler shift. Dectection of such a correlation in a sample of bursts would definitively confirm the rotational modulation model in our opinion. We will present calculations of mass - radius constraints for 4U 1636-54 and 4U 1728-34 including the effects of light bending, rotation and angle dependent emission, based on the observed properties of burst oscillations as well as spectroscopy of Eddington limited bursts in a subsequent paper. Several models for the kilohertz QPO seen in 13 LMXB systems (see \\markcite{vdk}van der Klis 1997 for a recent review) invoke some sort of beat-frequency interpretation for the twin kHz peaks seen in many of the sources (see \\markcite{MLP}Miller, Lamb \\& Psaltis 1997; \\markcite{S96}Strohmayer {\\it et al.} 1996). So far only in 4U 1728-34 does the frequency difference between the twin kHz peaks match the frequency observed during X-ray bursts (see \\markcite{S96}Strohmayer {\\it et al.} 1996). In two other sources (KS 1731-26 and 4U 1636-54) the separation of the twin kHz peaks is closer to 1/2 the frequency of oscillations observed during bursts (see \\markcite{WV}Wijnands \\& van der Klis 1997; and \\markcite{Z96}Zhang {\\it et al.} 1996). For example, \\markcite{W97}Wijnands {\\it et al.} (1997) report a frequency difference for the twin kHz QPO in 4U1636-54 of $276 \\pm 10$ Hz. The effort to reconcile these observations with a beat-frequency interpretation has led to speculation that the oscillation frequency observed during bursts may sometimes be twice the spin frequency of the star, although in 4U1636-54 the difference frequency appears to be a bit less than 1/2 the burst oscillation frequency, and in some Z sources the frequency difference is not constant (\\markcite{van97}van der Klis {\\it et al.} 1997). If this scenario is correct it implies the existence of two antipodal spots on the neutron star surface during X-ray bursts. In addition to the daunting requirement of initiating the thermonuclear flash nearly simultaneously on opposite sides of the neutron star, the observation of large oscillation amplitudes shortly after burst onset in 4U1636-54 places severe constraints on the two hotspot scenario. To see this one can ask the following question. What maximum amplitude can be produced by a star with antipodal hotspots? For two antipodal spots light bending strongly constrains the amplitudes that can be achieved (see \\markcite{PFC}Pechenick, Ftaclas \\& Cohen 1983). Calculations using the Schwarzchild spacetime and isotropic emission from the stellar surface indicate that even a neutron star with an implausibly small compactness of $M/R = 0.1$, recall that rotational modulation amplitude increases with decreasing compactness, can only achieve a maximum amplitude of about 30 \\%, whereas we measured an amplitude of 75\\% from the burst described here. We note that with $M/R = 0.1$ a 1.4 $M_{sun}$ neutron star would have a radius of 21 km. This is far stiffer than any neutron star equation of state that we are familiar with. Rapid rotation could in principle modify this result, but at the modest inferred spin period of 290.5 Hz for the two spots to produce the observed 581 Hz frequency one would still require implausibly large neutron star radii to make a significant rotational correction to the amplitude. Another process which can increase the amplitude is beaming of radiation at the stellar surface. For example, to the extent that electron scattering is the dominant opacity process in neutron star atmospheres then one should expect a specific intensity distribution which approximates that from a grey atmosphere. Such a distribution is proportional to $\\cos\\delta + 2/3$, where $\\delta$ is the angle from the normal to the stellar surface (see \\markcite{Mihal}Mihalas 1978), so this will modestly increase the rotational modulation amplitude. In addition, it is likely that the emergent spectrum will also be a weak function of $\\delta$ (\\markcite{ML97}Miller \\& Lamb 1997). With more detailed modelling of the emission from hotspots of finite size it will be possible to place strong constraints on the antipodal hotspot interpretation for the burst oscillations. Finally, if further analysis such as pulse phase spectroscopy continues to support the interpretation of the burst oscillation frequency as the neutron star spin frequency in 4U1636-54, then kHz QPO frequency separations near 1/2 the spin frequency would suggest that the beat frequency interpretation may be untenable." }, "9803/astro-ph9803096_arXiv.txt": { "abstract": "{ To examine the evidence for hierarchical evolution on mass scales of $\\sim 10^{13}$-$10^{14} \\Mdot$, we apply a statistic that measures correlations between galaxy velocity and projected position (Dressler \\& Shectman 1988) to data for six poor groups of galaxies, HCG 42, HCG 62, NGC 533, NGC 2563, NGC 5129, and NGC 741. Each group has more than 30 identified members (Zabludoff \\& Mulchaey 1998ab). The statistic is sensitive to clumps of galaxies on the sky whose mean velocity and velocity dispersion deviate from the kinematics of the group as a whole. The kinematics of galaxies within $\\sim 0.1$\\lith\\inv\\ Mpc of the group center do not deviate from the global values, supporting our earlier claim that the group cores are close to virialization or virialized. We detect significant substructure (at $\\geq 99.9$\\% confidence) in the two groups with the most confirmed members, HCG 62 and NGC 741, that is attributable mostly to a subgroup lying $\\sim 0.3$-0.4\\lith\\inv\\ Mpc outside of the core. We conclude that at least some poor groups, like rich clusters, are evolving via the accretion of smaller structures from the field. With larger poor group surveys, the incidence of such accretion and the distribution of subgroup masses are potential constraints of cosmological models on mass scales of $\\simless 10^{13}$-$10^{14} \\Mdot$ and on physical scales of $\\simless 0.5$\\lith\\inv\\ Mpc. \\vskip 0.5cm \\noindent{\\it Subject headings}: galaxies: clustering --- cosmology: large-scale structure of Universe } \\vfill\\eject ", "introduction": "The evolution of structure on different mass scales is one of the outstanding issues in cosmology. For example, although galaxy clusters of $\\sim 10^{15} \\Mdot$ (including Virgo (Binggeli 1993; Bohringer \\etal 1994), Coma (Mellier \\etal 1988; Briel \\etal 1992; White \\etal 1993), and Abell 754 (Zabludoff \\& Zaritsky 1995)) are clearly evolving from the accretion of smaller groups, it is uncertain whether poor groups of $\\sim 10^{13}$-$10^{14} \\Mdot$ also evolve hierarchically. There is some indirect evidence that the evolution of poor groups is similar to rich clusters; the galaxies and hot gas in poor groups follow the same relationships found among the X-ray temperature, X-ray luminosity, and galaxy velocity dispersion for rich clusters (Mulchaey \\& Zabludoff 1998, hereafter MZ98). Historically, however, the number of known poor group members has been too small to examine individual groups for direct evidence of hierarchical evolution. Multi-object spectroscopy now makes it possible to obtain ``cluster-size'' samples of galaxies in poor groups and to identify substructure, if it exists, in the same manner as for rich clusters. Substructure in clusters was not detected until the number of spectroscopically-confirmed cluster members exceeded $\\sim 30$-50 galaxies. Recent poor group surveys have reached this membership level (Zabludoff \\& Mulchaey 1998ab; hereafter ZM98a and b). The discovery of substructure in poor groups would provide new cosmological constraints by establishing that hierarchical evolution is occurring on mass scales of $\\sim 10^{13}$-$10^{14} \\Mdot$ and on physical scales of $\\sim 0.5$\\lith\\inv\\ Mpc. The detection of substructure would also support the picture in which at least some poor groups evolve as low-mass analogs to rich clusters. In this Letter, we describe the results from applying a substructure statistic (Dressler \\& Shectman 1988; hereafter DS88) to the six best sampled poor groups in ZM98ab, the most detailed spectroscopic survey of poor groups to date. ", "conclusions": "To search for substructure in the six poor groups, we use the method applied by DS88 to rich clusters. This test identifies a fixed number of nearest neighbors on the sky around each galaxy, calculates the local mean velocity and velocity dispersion of the subsample, and compares these values with the mean velocity and velocity dispersion of the entire group. The kinematic deviations of the subsamples from the global values are summed. This sum is larger for a group with a kinematically distinct subgroup than for a similar group without substructure. For each galaxy $i$, the deviation of its nearest projected neighbors from the kinematics of the group as a whole is defined as ${\\delta_i} \\equiv (n^{1/2}/ \\sigma_r) \\lbrack (\\upsilon_{loc} - \\overline \\upsilon)^2 + (\\sigma_{r,loc} - \\sigma_r)^2 \\rbrack ^{1/2}$, where $\\overline \\upsilon$ is the mean velocity for the group, $\\sigma_r$ is the group velocity dispersion, and $n$ is the number of nearest neighbors (including the galaxy) used to determine the local mean velocity $\\overline \\upsilon_{loc}$ and local velocity dispersion $\\sigma_{r,loc}$. The total deviation for the group is defined as the sum of the local deviations, $\\Delta \\equiv \\sum |\\delta_i|$ for all $i \\leq N_{grp}$, the number of group members. As pointed out in DS88, the $\\Delta$ statistic is similar to the $\\chi^2$ statistic, except that the $\\delta_i$'s are not squared before summation in order to reduce the contributions of the largest, rarest deviations. If the galaxy velocity distribution of the group is close to Gaussian, and the local variations are only random fluctuations, $\\Delta \\simeq N_{grp}$. To calculate $\\delta$, we choose $n = 11$ (as in DS88). This choice allows robust determinations of $\\overline \\upsilon_{r,loc}$ and $\\sigma_{r,loc}$. Silverman (1986) argues that using $n \\sim {N_{grp}}^{1/2}$ nearest neighbors (= 6-8 for these groups) maximizes the sensitivity of such a test to small scale structures while reducing its sensitivity to fluctuations within the Poisson noise (also see Bird 1994b). To check the robustness of the $n=11$ assumption, we compare the results below with those for $n=6$. The conclusions drawn from the $n=6$ and $n=11$ cases are the same. Calibration of the $\\Delta$ statistic for each group is required because 1) the $\\delta_i$'s are not statstically independent and 2) the velocity distribution may not be intrisically Gaussian even if there are no subgroups ({\\it e.g.}, the group members may follow predominantly circular or radial orbits). We determine the significance of the observed $\\Delta$ by comparing it with the results of 1000 Monte Carlo trials in which galaxy velocities are drawn randomly from the observed distribution and assigned to galaxy positions. This scrambling technique effectively destroys any substructure (DS88). If the probability is low that a group without substructure has a $\\Delta$ value at least as large as that observed, then we consider the substructure detection significant. In two groups, HCG 62 and NGC 741, the observed value of $\\Delta$ is significant at the $\\geq 99.9\\%$ confidence level. Such high $\\Delta$ values might arise from substructure, but also could result from smooth variations in the group's velocity field ({\\it e.g.}, rotation or velocity shear (Malumuth \\etal 1992)) and/or from a dependence of $\\sigma_r$ on radius (Bird 1994b). ZM98a show that $\\sigma_r$ is constant out to radii of $\\sim 0.5$\\lith\\inv\\ Mpc in the combined velocity dispersion profile for the sample groups. To determine whether substructure is in fact responsible for the high $\\Delta$ values in HCG 62 and NGC 741, we examine the local deviation $\\delta$ for each group member. A concentration of large $\\delta$ values on the sky indicates a kinematically distinct subgroup. Figure 1 shows the projected spatial distribution of group members for each group (top panel). The second panel shows this distribution with the radii of the circles weighted by $e^\\delta$ (as in DS88). Because each point is not statistically independent, a few very deviant galaxies can boost the $\\delta$ values for a large number of nearby points. Therefore, a visual comparison with the Monte Carlo simulations is required to assess the significance of any structures. The third panel shows the results of the Monte Carlo trial (out of 1000) with the largest $\\Delta$ value, or greatest total deviation. The results of the trial with the median $\\Delta$ value are in the bottom panel. The significance of the seven large clustered circles to the northeast of HCG 62 and five large clustered circles to the south of NGC 741 is high. In each case, the large $\\delta$ values show that a subgroup not in equilibrium with the global group potential is the principal source of the significant $\\Delta$ value. Each of the two subgroups lies a projected distance of $\\sim 0.3$-0.4\\lith\\inv\\ Mpc outside of the group core. On the other hand, the clustering of small circles within $\\sim 0.1$\\lith\\inv\\ Mpc of all of the group centers indicates that the core mean velocity and velocity dispersion are similar to the global values for the group. This result suggests that the group cores are close to virialization or virialized and is consistent with the conclusions from our earlier studies of group dynamics (ZM98a, MZ98)." }, "9803/astro-ph9803089_arXiv.txt": { "abstract": "Aperiodic variability and Quasi Periodic Oscillations (QPOs) are observed from accretion disks orbiting white dwarfs, neutron stars, and black holes, suggesting that the flow is universally broken up into discrete blobs. We consider the interaction of these blobs with the magnetic field of a compact, accreting star, where diamagnetic blobs suffer a drag. We show that when the magnetic moment is not aligned with the spin axis, the resulting force is pulsed, and this can lead to resonance with the oscillation of the blobs around the equatorial plane; a resonance condition where energy is effectively pumped into non--equatorial motions is then derived. We show that the same resonance condition applies for the quadrupolar component of the magnetic field. We discuss the conditions of applicability of this result, showing that they are quite wide. We also show that realistic complications, such as chaotic magnetic fields, buoyancy, radiation pressure, evaporation, Kelvin--Helmholtz instability, and shear stresses due to differential rotation do not affect our results. In accreting neutron stars with millisecond periods, we show that this instability leads to Lense--Thirring precession of the blobs, and that damping by viscosity can be neglected. ", "introduction": "The interaction between the magnetic field anchored to rotating bodies and matter orbiting around them plays an important role in a variety of astrophysical situations, ranging from the Jupiter-Io system (Stern and Ness 1982) to the Jovian ring (Burns \\etal\\/ 1985), to protoplanetary disks around young magnetic stars (Bodenheimer 1995) and accreting degenerate stellar remnants in binary systems, such as white dwarfs in intermediate polars or neutron stars in X-ray binaries (Frank, King and Raine 1995). Straightforward applications of the laws of electrodynamics are possible in cases, such as the Jovian ring, in which the orbiting matter is made of solid, dust particles. Due to a variety of poorly known MHD and plasma effects, modelling this interaction in the case of gaseous disks that orbit different classes of magnetic stars is far more uncertain and difficult. In particular, theoretical descriptions of viscous accretion disks around magnetic rotating stars have been largely based on a number of simplifying assumptions. Two of these are especially relevant to the present work: (a) that the magnetic field (assumed dipolar) and rotation axes of the star are coaligned; (b) that the disk is continuous, smooth, azimuthally symmetric and lies in the equatorial plane of the rotating star. For example Ghosh and Lamb (1979) adopted these assumptions and introduced an {\\it ad hoc} effective diffusivity of the disk plasma to derive a stationary disk solution, in which the star magnetic field lines thread the disk and slip across its material. Models of this kind proved very useful for evaluating the basic properties of the disk-magnetic field interaction, such as the size of the magnetosphere and the balance of the material and non-material torques. It has always been recognised that the coalignement of the magnetic field and rotation axes of the star is an unrealistic hypothesis as a finite magnetic colatitude is required for the generation of the periodic signals at the star spin frequency that are often observed in these systems. On the other hand, the pronounced aperiodic (or, in some cases, quasi-periodic) variability that is frequently detected on a variety of timescales ranging from days to milliseconds in disk accreting compact stars of all classes (intermediate polars: King and Lasota 1990; accreting neutron stars and black holes: van der Klis 1995, van der Klis 1997) testify that the disk cannot be regarded as smooth, continuous and azimuthally symmetric. In a number of cases the quasi-periodic timing signature of blobs orbiting over a limited range of radii close to the compact star is clearly seen. A highly inhomogeneous and clumpy disk, possibly comprising two distinct and coexisting phases (hot and cold), is also envisaged as the end product of the instabilities predicted by current models of viscous disks (e.g the so-called secular and viscous instabilities of radiation-pressure dominated $\\alpha$-disks, Lightman and Eardley 1974, Shakura and Sunyaev 1976, or the magnetic amplification and buoyancy of flux tubes in dynamo-driven disks, Vishniac and Diamond 1992), as discussed by Krolik (1998). In all accreting objects endowed with a strong magnetic field, the presence of inhomogeneities and/or blobs that can be regarded as discrete entities suggests that a novel mode of interaction between the compact star and the disk is possible. This occurs because individual blobs are most likely strongly diamagnetic (see for instance King 1993, Wynn and King 1995 and references therein); when moving through a magnetic field, strong surface currents develop on the blob the main effect of which is the generation of a drag opposite to the component of the blob velocity perpendicular to the field (Drell, Foley and Ruderman 1965). The acceleration acting on each blob is thus \\begin{equation} \\vec{a} = -\\frac{\\vec{v}_\\perp^{(rel)}}{t_d}\\;; \\end{equation} the drag time--scale $t_d$ is given by \\begin{equation} t_d = \\frac{c_A m}{B^2 l^2}\\;, \\end{equation} where $c_A$ is the Alfv\\`en speed in the magnetic field $B$, and $m$ and $l$ are the mass and characteristic radius of the blob. Here $\\vec{v}^{(rel)}$ is the relative velocity between the blob and a magnetic field line. It is easy to see (Fig.1) that this acceleration has a component along the disk axis which tends to lift the blob off the equatorial plane where it is, at least initially, lying. It is the purpose of this paper to study this dynamical interaction, and to show how this alters the conventional view of accretion disks onto magnetized compact stars. In order to bring out the physical meaning of the instability most clearly, we shall at first idealize these plasma blobs as point masses. In the next section, it will be shown that this interaction leads to the lifting of blobs off the equatorial plane, at a resonance radius; the conditions under which this result applies are discussed in Section 3. In Section 4, we relax the hypothesis that blobs are point masses, and establish that a variety of effects, all related to the blobs having a finite size, do not modify our results. As the only concrete application of this instability, we discuss in Section 5 the generation of modulation of the X--ray flux in Low Mass X--ray Binaries (LMXBs) exhibiting millisecond QPOs at frequencies comparable to the single--particle Lense--Thirring (1918) precession frequency. The last section summarizes the results. ", "conclusions": "The processes described above are generic: they apply to every accretion disk broken up into discrete units, surrounding a magnetized non--aligned rotator provided the ordering of time--scales (Eq. \\ref{ordering}) holds and the blobs have density contrasts $(\\rho_b-\\rho_d)/ \\rho_d \\ga 1$. The instability is independent of the detailed diamagnetic properties of the blobs (here enshrined in the parameter $t_d$ which is just modulated at the relative frequency $\\omega_K-\\omega_s$), of their masses and densities, their interactions with radiation emitted by the accreting source, and the exact form of the viscosity. Provided the hierarchy of timescales $t_K \\ll t_d \\ll t_v$ (Eq. \\ref{ordering}) is established and blobs have non--negligible density contrasts, it should apply to both accreting white dwarfs and neutron stars. The existence of resonances is essentially due to the fact that a spatially periodic magnetic field, such as that due to a single multipole, will produce a temporally--variable electromagnetic force on a single blob, as it passes through the various frequency components of the field. Minor corrections to the exact locations of the resonances might derive from effects we opted not to consider, such as an axially offset field, or a rotation rate different for different multipoles, as it happens to the non--dipolar part of the magnetic field of the Earth. Another phenomenon we neglected to investigate, and which might generate further resonances, is the non--linear coupling of radial, latitudinal and longitudinal motions. It should however be borne in mind that the exact locations of resonances might be irrelevant since blobs are expected to have a finite size. The evolution of the accretion disk after the resonance of Eq. \\ref{resonance} is very difficult to predict; it seems clear enough that blobs will tend to move closer to the magnetic equatorial plane, and oscillate around it, but the further evolution is uncertain because the details depend on a very poorly known quantity, the $z$--axis viscosity. This is the reason why we did not push the analysis performed in this paper into the non--linear regime. It seems quite clear that the disk must puff--up, but whether, and when, viscosity will manage to bring it back to the equatorial plane remains an open question. Very interesting consequences of the above--discussed instability occur in a specific class of accretors, \\ie\\/ neutron stars exhibiting millisecond QPOs. Blobs that have acquired $z$~axis motions will be acted upon by the torque which causes Lense--Thirring (1918, LT) precession. We have shown here that, in both Atoll and Z--type sources, viscosity is surely unable to bring blobs back down to the equatorial plane. For motions in the plane, the net effect of the instability is to induce epicyclic motions, or, given that in $1/r$ potentials all orbits close, to perturb the blobs' motions into ellipses. Then periastron precession induced by general relativistic effects ought to ensue. A related mechanism for the generation of off--plane motions has been proposed for the Jovian dust ring (Burns \\etal\\/ 1985). It differs from the present one in that the individual units are not blobs of plasma, but single dust grains with nonzero electric charge, which are then acted upon by the normal Lorenz force $q \\vec{v}^{(rel)}\\wedge\\vec{B}/c$. It is easy to see that this differs from our case because the forces differ: in our case, the acceleration (Eq. 1) is in the plane of the two vectors $\\vec{v}^{(rel)}$ and $\\vec{B}$, while in the Jovian case the force is perpendicular to this plane. This difference has as a consequence that, in the Jovian case, each magnetic field multipole has its own, single resonance, while in our case the two multipoles we explored have the same array of resonances, and the array is an infinite one. In short, in this paper we have shown that a simple interaction lifts blobs off the equatorial plane of a viscous accretion disk, provided the accreting object is not an aligned rotator. This occurs in a thin radial annulus, identified by Eq. \\ref{resonance}. This conclusion is rather general, being independent of the exact form of the drag time--scale, of the blobs' diamagnetic properties, and also of the assumed form for the disk viscosity. We have also shown why viscosity is unable to damp these motions (Section \\ref{lt}), and that this leads to modulation of the X--ray flux at the single particle precession frequency, or possibly twice as much. Thus, the proposed identification of some QPOs in the spectra of both Atoll and Z sources as due to LT--precession (Stella and Vietri 1998) is made more likely by the existence of a relevant mechanism exciting blobs' motions off the equatorial plane. Helpful comments from F.K. Lamb and J. Imamura are gratefully acknowledged. This work was begun while one of us (M.V.) was visiting the Institute for Advanced Study; its Director, John Bahcall, is thanked for the warm hospitality. The work of L.S. was partially supported through ASI grants." }, "9803/astro-ph9803040_arXiv.txt": { "abstract": "With the new generation of instruments for Cosmic Microwave Background (CMB) observations aiming at an accuracy level of a few percent in the measurement of the angular power spectrum of the anisotropies, the study of the contributions due to secondary effects has gained impetus. Furthermore, a reinvestigation of the main secondary effects is crucial in order to predict and quantify their effects on the CMB and the errors that they induce in the measurements. \\par In this paper, we investigate the contribution, to the CMB, of secondary anisotropies induced by the transverse motions of clusters of galaxies. This effect is similar to the Kaiser--Stebbins effect. In order to address this problem, we model the gravitational potential well of an individual structure using the Navarro, Frenk \\& White profile. We generalise the effect of one structure to a population of objects predicted using the Press-Schechter formalism. We simulate maps of these secondary fluctuations, compute the angular power spectrum and derive the average contributions for three cosmological models. We then investigate a simple method to separate this new contribution from the primary anisotropies and from the main secondary effect, the Sunyaev-Zel'dovich kinetic effect from the lensing clusters. \\par ", "introduction": "During the next decade, several experiments are planned to observe the Cosmic Microwave Background (CMB) and measure its temperature fluctuations (Planck surveyor, Map, Boomerang, ...). Their challenge is to measure the small scales anisotropies of the CMB (a few arcminutes up to ten degrees scale) with sensitivities better by a factor 10 than the COBE satellite (Smoot et al. 1992). These high sensitivity and resolution measurements will tightly constrain the value of the main cosmological parameters (Kamionkowski et al. 1994). However, the constraints can only be set if we are able to effectively measure the {\\it primary} temperature fluctuations. These fluctuations, present at recombination, give an insight into the early universe since they are directly related to the initial density perturbations which are the progenitors to the cosmic structures (galaxies and galaxies clusters) in the present universe; but which are first and foremost the relics of the very early initial conditions of the universe.\\\\ Between recombination and the present time, the CMB photons could have undergone various interactions with the matter and structures present along their lines of sight. Some of these interactions can induce additional temperature fluctuations called, {\\it secondary} anisotropies because they are generated after the recombination. Along a line of sight, one measures temperature fluctuations which are the superposition of the {\\it primary } and {\\it secondary} anisotropies. As a result, and in the context of the future CMB experiments, accurate analysis of the data will be needed in order to account for the foreground contributions due to the secondary fluctuations. Photon--matter interactions between recombination and the present time are due to the presence of ionised matter or to variations of the gravitational potential wells along the lines of sight. \\par The CMB photons interact with the ionised matter mainly through Compton interactions. In fact, after recombination the universe could have been re-ionised globally or locally. Global early re-ionisation has been widely studied (see Dodelson \\& Jubas 1995 for a recent review and references therein). Its main effect is to either smooth or wipe out some of the primary anisotropies; but the interactions of the photons with the matter in a fully ionised universe can also give rise to secondary anisotropies through the Vishniac effect (Vishniac 1987). This second order effect has maximum amplitudes for a very early re-ionisation. The case of a late inhomogeneous re-ionisation and its imprints on the CMB fluctuations has been investigated (Aghanim et al. 1996) and found to be rather important. In this case, the secondary anisotropies are due to the bulk motion of ionised clouds with respect to the CMB frame. When the re-ionisation is localised in hot ionised intra-cluster media the photons interact with the free electrons. The inverse Compton scattering between photons and electrons leads to the so-called Sunyaev-Zel'dovich (hereafter SZ) effect (Sunyaev \\& Zel'dovich1972, 1980). The Compton distortion due to the motion of the electrons in the gas is called the thermal SZ effect. The kinetic SZ effect is a Doppler distortion due to the peculiar bulk motion of the cluster with respect to the Hubble flow. The SZ thermal effect has the unique property of depressing the CMB brightness in the Rayleigh-Jeans region and increasing its brightness above a frequency of about 219 GHz. This frequency dependence makes it rather easy to observe and separate from the kinetic SZ effect. In fact, the latter has a black body spectrum which makes the spectral confusion between kinetic SZ and primary fluctuations a serious problem. The SZ effect has been widely studied for individual clusters and for populations of clusters. For full reviews on the subject we refer the reader to two major articles: Rephaeli 1995 and Birkinshaw 1997. These investigations have clearly shown that the SZ effect in clusters of galaxies provides a powerful tool for cosmology through measurements of the Hubble constant, the radial peculiar velocity of clusters and consequently the large scale velocity fields. \\par Besides the interactions with the ionised matter, some secondary effects arise when the CMB photons traverse a varying gravitational potential well. In fact, if the gravitational potential well crossed by the photons evolves between the time they enter the well and the time they leave it, the delay between entrance and exit is equivalent to a shift in frequency, which induces a temperature anisotropy on the CMB. This effect was first studied by Rees \\& Sciama (1968) for a potential well growing under its own gravity. Numerous authors have investigated the potential variations due to collapsing objects and their effect on the CMB (Kaiser 1982, Nottale 1984, Martinez-Gonz\\'alez, Sanz \\& Silk 1990, Seljak 1996). Similarly, a gravitational potential well moving across the line of sight is equivalent to a varying potential and will thus imprint secondary fluctuations on the CMB. This effect was first studied for one cluster of galaxies by Birkinshaw \\& Gull (1983) (Sect. 2). Kaiser \\& Stebbins (1984) and Bouchet, Bennett \\& Stebbins (1988) investigated a similar effect for moving cosmic strings. Recent work (Tuluie \\& Laguna 1995, Tuluie, Laguna \\& Anninos 1996) based on N-body simulations has pointed out this effect in a study of the effect of varying potential on rather large angular scales ($\\simeq 1^{\\circ}$). A discussion of some of these results and a comparison with ours will follow in the next sections. \\par In this paper, following the formalism of Birkinshaw \\& Gull (1983) and Birkinshaw (1989), we investigate the contribution of secondary anisotropies due to a population of collapsed objects moving across the line of sight, these objects range from small groups to rich clusters in scale ($10^{13}$ to $10^{15}$ $M_{\\sun}$). In section 2., we first study in detail the case of a unique collapsed structure. We use a structure model to compute in particular the deflection angle and derive the spatial signature of the moving lens effect. We then account (Section 3.) for the contribution, to the primordial cosmological signal, of the whole population of collapsed objects using predicted counts and we simulate maps of these secondary anisotropies. In section 4., we analyse the simulated maps and present our results. We give our conclusions in section 5. ", "conclusions": "In our work, we investigate the secondary fluctuations induced by moving lenses with masses ranging from those of groups of galaxies to those of clusters of galaxies in a simple way, based on predicted structure counts and simulated maps. This method allows us to explore a rather wide range of scales ($>10$ arcseconds) in various cosmological models. The analysis, in terms of angular power spectra, show the scales for which the primary fluctuations are dominant (Fig. \\ref{fig:pspec}). In the standard and lambda CDM models, the primary anisotropies are dominant respectively for scales $l<4000$ and $l<4500$ whereas in the Open CDM model they are dominant for $l<6000$. In practice, it is thus impossible to detect the secondary anisotropies due to moving lenses in the open model. The standard CDM model shows the smallest cut-off scale with an intermediate SZ kinetic pollution, compared to the other two models. It is therefore the ``best case'' framework for making an analysis and predicting the detection of fluctuations and the contributions that they induce. One must keep in mind that the results quoted in this particular case represent the ``best'' results we get from the analysis. \\par The results of our analysis are obtained under the assumption of a universe that never re-ionises, which is of course not the case. The re-ionisation, if it is homogeneous, is supposed to somewhat ease the task of extraction of the pattern. In fact, its main effect is to damp the angular power spectrum of the primary anisotropies on small scales, shifting the cut-off towards larger scales. In this case, the effect of moving lenses dominates over the CMB fluctuations, and the SZ kinetic is not as high as it is on very small scales. However, if the re-ionisation is late and inhomogeneous, it generates additional SZ kinetic-type secondary fluctuations (Aghanim et al. 1996) without damping the power spectrum by more than a few percent. Here, the re-ionisation might worsen the analysis at small scales. In any case, there could be some other additional secondary fluctuations principally due to the Vishniac effect, that arise in a re-ionised universe. Our work thus gives a ``best case'' configuration of the problem, with all other effects tending to worsen the situation. \\par We found that the secondary fluctuations induced by the moving gravitational lenses can be as high as $1.5\\,10^{-5}$; with {\\it rms} contributions of about 5 to $3.\\,10^{-7}$ in the three cosmological models. Even if the moving lens fluctuations have a particular dipolar pattern and even if they are ``perfectly'' located through their SZ thermal effect, the detection of the moving lens effect and its separation from the SZ kinetic and primary fluctuations are very difficult because of the very high level of confusion, on the scales of interest, with the point--like SZ kinetic anisotropies and because of spectral confusion. \\par We nevertheless analysed the simulated maps using an adap\\-ted wavelet technique in order to extract the moving lens fluctuations. We conclude that {\\it the contribution of the secondary anisotropies due to the moving lenses is thus negligible whatever the cosmological model. Therefore it will not affect the future CMB measurements except as a background contribution. We have highlighted the fact that the moving lens fluctuations have a very significant spatial signature but we did not succeed in separating this contribution from the other signals.}" }, "9803/astro-ph9803276_arXiv.txt": { "abstract": "Images recorded through broad ($J, H, K$), and narrow (CO, and $2.2\\mu$m continuum) band filters are used to investigate the photometric properties of bright ($K \\leq 13.5$) stars in a $6 \\times 6$ arcmin field centered on the SgrA complex. The giant branch ridgelines in the $(K, J-K)$ and $(K, H-K)$ color-magnitude diagrams are well matched by the Baade's Window (BW) M giant sequence if the mean extinction is A$_K \\sim 2.8$ mag. Extinction measurements for individual stars are estimated using the M$_K$ versus infrared color relations defined by M giants in BW, and the majority of stars have A$_K$ between 2.0 and 3.5 mag. The extinction is locally high in the SgrA complex, where A$_K \\sim 3.1$ mag. Reddening-corrected CO indices, CO$_o$, are derived for over 1300 stars with $J, H$, and $K$ brightnesses, and over 5300 stars with $H$ and $K$ brightnesses. The distribution of CO$_o$ values for stars with $K_o$ between 11.25 and 7.25 can be reproduced using the M$_K -$ CO$_o$ relation defined by M giants in BW. The data thus suggest that the most metal-rich giants in the central regions of the bulge and in BW have similar photometric properties and $2.3\\mu$m CO strengths. Hence, it appears that the central region of the bulge does not contain a population of stars that are significantly more metal-rich than what is seen in BW. \\vspace{0.3cm} \\noindent{Key words: Galaxy: center -- stars: abundances -- stars: late-type} ", "introduction": "The stars in the central regions of the Galaxy have been the target of numerous photometric and spectroscopic investigations. The brightest, best-studied, objects belong to a young population (e.g. reviews by Blum, Sellgren, \\& DePoy 1996a, and Morris \\& Serabyn 1996) that dominates the near-infrared light within 0.1 -- 0.2 parsec ($\\sim 2.5 - 5.0$ arcsec) of SgrA* (Saha, Bicknell, \\& McGregor 1996), and is concentrated within the central $\\sim 1$ parsec (Allen 1994). However, there is also a large population of older stars near the Galactic Center (GC) belonging to the inner bulge, and it is only recently that efforts have been made to probe the nature of these objects. Minniti {\\it et al.} (1995) reviewed metallicity estimates for a number of bulge fields, and a least squares fit to these data indicates that $\\Delta$[Fe/H]/$\\Delta$log(r) $= -1.5 \\pm 0.4$ (Davidge 1997). The presence of a metallicity gradient suggests that the material from which bulge stars formed experienced dissipation, so it might be anticipated that the central regions of the bulge will contain the most metal-rich stars in the Galaxy. While the fields considered by Minniti {\\it et al.} are at relatively large distances from the GC, and hence do not monitor trends at small radii, observations of other galaxies indicate that population gradients can extend into the central regions of bulges. This is clearly evident in spectroscopic studies of M31 (e.g. Davidge 1997), a galaxy that shares some morphological similarities with the Milky-Way (e.g. Blanco \\& Terndrup 1990). It is not clear from the existing observational data if the bright stellar contents in the inner bulge and Baade's Window (BW), a field that is dominated by stars roughly 0.5 kpc from the GC, are similar. The $K$ luminosity function (LF) of moderately faint stars within an arcmin of SgrA* has a power-law exponent similar to that seen in BW (Blum {\\it et al.} 1996a; Davidge {\\it et al.} 1997). However, the significance of this result is low, as the LFs of bright giant branch stars are insensitive to metallicity (e.g. Bergbusch \\& VandenBerg 1992). The brightest inner bulge stars, which are evolving on the asymptotic giant branch (AGB), have $2\\mu$m spectroscopic properties reminiscent of bright giants in BW, although detailed measurements reveal that for a given equivalent width of near-infrared Na and Ca absorption, the inner bulge stars have deeper CO bands than giants in BW (Blum, Sellgren, \\& DePoy 1996b). While suggestive of differences in chemical composition, it should be recalled that these objects are the brightest, most highly evolved members of the bulge, and Na can be affected by mixing (e.g. Kraft 1994). Hence, the spectroscopic properties of these bright red giants may not be representative of fainter objects. As the strongest features in the near-infrared spectra of cool evolved stars, the $2.3\\mu$m first-overtone CO bands provide an important means of probing the stellar content of the inner bulge. Unfortunately, the crowded nature of the inner bulge, coupled with the low multiplex advantage offered by the current generation of cryogenically-cooled spectrographs, most of which use a single long slit, makes a $2\\mu$m spectroscopic survey of a large sample of moderately faint objects a difficult task at present. Narrow-band imaging, using filters such as those described by Frogel {\\it et al.} (1978), provides a highly efficient alternate means of measuring the strength of CO absorption in a large number of objects. In the current paper $J, H, K$, CO and $2.2\\mu$m continuum measurements of moderately faint ($K \\leq 13.5$) stars are used to measure the strength of CO absorption in stars within 3 arcmin of the GC. To the best of our knowledge, this is the largest survey of stellar content in the central regions of the bulge conducted to date. The observations and reduction techniques are described in \\S 2, while the photometric measurements are discussed in \\S 3. In \\S 4 the line-of-sight extinction to these sources is estimated by assuming that they follow the same M$_K -$ color relations as M giants in BW. Reddening-corrected CO indices are derived, and the distribution of CO indices is compared with that predicted if stars in the inner bulge and BW follow similar M$_K -$ CO relations. A summary and brief discussion of the results follows in \\S 5. ", "conclusions": "Moderately deep near-infrared images have been used to probe the bright ($K \\leq 13.5$) stellar content in the central $6 \\times 6$ arcmin field of the Galaxy. The ridgeline of the bulge giant branch on the $(K, J-K)$ and $(K, H-K)$ CMDs is well matched by the BW M giant sequence, reddened according to the Rieke \\& Lebofsky (1985) extinction curve. This similarity in photometric properties suggests that the extinctions to individual stars in the inner bulge can be estimated by adopting the M$_K -$ color relations defined by M giants in BW. The mean extinction outside of the SgrA complex, where A$_K = 3.1$ mag, is A$_K = 2.8$ mag. The extinction estimates for individual stars have been used to generate reddening-corrected CO indices, and the histogram distribution of CO$_o$ values can be reproduced using the M$_K -$ CO$_o$ relation defined by M giants in BW. Therefore, M giants near the GC and in BW have similar $2.3\\mu$m CO strengths. A potential source of systematic error in the procedure used to compute A$_K$ is that a single set of M$_K -$ color relations have been used, and no attempt has been made to allow for a dispersion in the metallicities of giants in the central regions of the bulge. There is a selection effect for magnitude-limited samples, in the sense that the brightest stars in a field containing an old composite population are likely the most metal-rich, so the current data likely do not sample the full range of metallicities near the GC. In any event, the CO$_o$ distribution does not change substantially when M$_K -$ color relations defined by red giants in 47 Tuc ([Fe/H] $\\sim -0.7$) are used to estimate extinction, indicating that the main conclusions of this paper are insensitive to the adopted intrinsic colors of GC giants. Another source of systematic error is that the A$_K$ values derived in \\S 4 require measurements in $J, H$, and $K$. This broad wavelength coverage introduces a bias against heavily reddened objects, which are faint, and hence may not be detected, in $J$. In fact, there are regions where the extinction is so high that stars are not detected in $K$. The tendency to miss the most heavily reddened stars skews the A$_K$ distribution to lower values. One way to reduce, but not entirely remove, this bias is to estimate extinction using only $H-K$ colors, for which the number of objects is over $3 \\times$ greater than those with $J$ measurements. Following the procedure described earlier, $(H-K)_0$ was assigned to each star using the $M_K - (H-K)$ relation for BW M giants, and the results are shown in the lower panels of Figures 7 -- 12, while the radial distribution of CO$_o$ values is plotted in Figure 14. It is evident from the lower panel of Figure 8 that a number of sources with relatively high A$_K$ are added to the sample when measurements in $J$ are not required. Nevertheless, the impact on the CO$_o$ distribution is negligible, indicating that stars with higher than average obscuration near the GC have $2.3\\mu$m CO strengths that are similar to less heavily reddened stars. It should be emphasized that the $2.3\\mu$m CO bands provide only one diagnostic of chemical composition. Indeed, studies of the radial behaviour of the $2.3\\mu$m CO bands in the bulges of other galaxies reveal weak or non-existant gradients (e.g. Frogel {\\it et al.} 1978), even though many of these systems show line strength gradients at optical wavelengths. When interpreting this ostensibly contradictory result it should be recalled that the CO measurements are dominated by the brightest red stars which, in old populations, will be those that are the most metal-rich. If the most metal-rich population in the bulges of other galaxies follows a single M$_K -$ CO relation with no radial dependence, as appears to be the case in the inner regions of the Galactic bulge, then this would help to explain why the wide-aperture CO measurements of other systems do not show gradients. There are indications that the abundances of some species in the spectra of metal-rich giants may vary with distance from the GC. In particular, the infrared spectra obtained by Blum {\\it et al.} (1996b) indicate that $2.3\\mu$m CO absorption in GC giants is {\\it stronger} than in BW stars having the same $2\\mu$m Na and Ca absorption line strengths. Therefore, given that the CO line strengths in giants near the GC are similar to those in BW, then the Blum et al. measurements are suggestive of radial changes in [Na/Fe] and [Ca/Fe] among bulge stars. While this paper has concentrated on bulge stars, a modest disk population is also evident in the CMDs, and these data can be used to estimate the contribution disk objects make to the near-infrared light output near the GC. Objects with $(J-K) \\leq 2.5$ and brightnesses between $K \\sim 7.5$ (the approximate saturation limit of the CTIO data) and $K \\sim 12$ (the faintest point at which the CTIO $K, J-K$ CMD is complete over a broad range of colors) account for 2.6\\% of the total light from resolved sources within 3 arcmin of the GC. The relative contribution made by disk stars would be much lower if dust did not obscure the bulge. Assuming that (1) the disk stars are not heavily reddened, and (2) stars near the GC are obscured by A$_K \\sim 2.8$ mag then, after correcting for this extinction, the contribution from disk stars drops to only 0.05\\% when $K_0 \\leq 8.5$. \\vspace{0.3cm} Sincere thanks are extended to the referee, Jay Frogel, for providing comments that greatly improved the paper. \\clearpage \\begin{table*} \\begin{center} \\begin{tabular}{cccr} \\tableline\\tableline IRS \\# & K$_{TD}$ & K$_{BSD}$ & $\\Delta$ \\\\ \\tableline 7 & 6.44 & 6.55 & $-0.11$ \\\\ 9 & 8.80 & 8.57 & 0.23 \\\\ 16NE & 8.56 & 9.01 & $-0.45$ \\\\ 28 & 9.41 & 9,36 & 0.05 \\\\ \\tableline \\end{tabular} \\end{center} \\caption{Comparison with $K$ measurements obtained by Blum {\\it et al.} (1996a)} \\end{table*} \\clearpage" }, "9803/astro-ph9803106_arXiv.txt": { "abstract": "Lithium is an excellent tracer of mixing in stars as it is destroyed (by nuclear reactions) at a temperature around $\\sim 2.5\\times 10^6$ K. The lithium destruction zone is typically located in the radiative region of a star. If the radiative regions are stable, the observed surface value of lithium should remain constant with time. However, comparison of the meteoritic and photospheric Li abundances in the Sun indicate that the surface abundance of Li in the Sun has been depleted by more than two orders of magnitude. This is not predicted by solar models and is a long standing problem. Observations of Li in open clusters indicate that Li depletion is occurring on the main sequence. Furthermore, there is now compelling observational evidence that a spread of lithium abundances is present in nearly identical stars. This suggests that some transport process is occurring in stellar radiative regions. Helioseismic inversions support this conclusion, for they suggest that standard solar models need to be modified below the base of the convection zone. There are a number of possible theoretical explanations for this transport process. The relation between Li abundances, rotation rates and the presence of a tidally locked companion along with the observed internal rotation in the Sun indicate that the mixing is most likely induced by rotation. The current status of non-standard (particularly rotational) stellar models which attempt to account for the lithium observations are reviewed. ", "introduction": "Li\\footnote{In this review I will use Li to represent $^7$Li, the isotope which is produced by big bang nucleosynthesis. $^7$Li accounts for $\\sim 93\\%$ of the total Li abundance in meteorites. Observers typically measure the total Li abundance, while theoretical models determine depletion factors for $^7$Li and $^6$Li The $^6$Li isotope is destroyed at much lower temperatures and $^7$Li. When making the comparison between the observations and theory, it is usually assumed that the $^6$Li contribution to the stellar Li content is neglible. However, see the discussion on halo stars (\\S \\ref{secthalo}) where observations of $^6$Li may be used to elucidate the mixing mechanism operating in these stars.} is a sensitive tracer of mixing in stellar radiative regions as it is easily destroyed at temperatures above $\\sim 2.5 \\times 10^6\\,$K. For solar type stars, the Li destruction region is located below the surface convection zone in standard models. As a consequence, standard stellar models predict that Li should not be depleted at the surface of solar type stars. This is a rather robust prediction of stellar evolution theory, which has been known for 40 years \\cite{schwar}. However, comparisons between the solar photospheric Li abundance and the Li abundance in meteorites show that the Sun has depleted a substantial amount of Li at its surface \\cite{green51}. The solar Li depletion problem has posed a challenge to stellar evolution theory for 40 years, and the solution to this puzzle is still open to debate. The Sun is unique in that helioseismic observations allow us to probe the interior structure and rotation of the Sun. These observations can put constraints on possible solutions to the solar Li depletion problem, but by themselves solar observations cannot uniquely determine the cause of solar Li depletion. Observations of stellar Li abundances allow one to study the Li depletion problem as a function of age, metallicity and stellar mass. As such, they provide a powerful test for mechanisms which attempt to explain the solar Li depletion. The discovery of a large dip in Li abundances around 6600 K in the Hyades \\cite{lidip} was not predicted by theorists, and remains a major challenge to theoretical stellar evolution models. There is increasing evidence that a dispersion in Li abundances exists among stars with similar ages, metallicities and masses (\\citeauthor{soder93} \\citeyear{soder93}; \\citeauthor{boes98} \\citeyear{boes98}). Such a dispersion suggests that another stellar property is important in determining the amount of Li which is depleted in stars. There is mounting observational evidence that rotation plays a key role in determining the amount of Li depletion in a star (\\citeauthor{hyadbin} \\citeyear{hyadbin}; \\citeauthor{jones97} \\citeyear{jones97}). In this review, I will discuss the relationship between mixing, rotation and Li abundances in stars. ", "conclusions": "} There is a wealth of data on Li abundances and rotation velocities in stars with a variety of ages, masses and metallicities. This data clearly indicates that Li depletion occurs on the main sequence for all stars with a surface convection zone. This is in direct contradiction with standard stellar evolution theory. A number of possible mechanisms which lead to extra Li depletion have been put forth. The dispersion in Li abundances at a given age, metallicity and temperature, the correlation between Li abundances and rotation velocities in Pleiades, the fact that tidally locked binary stars in the Hyades have an excess Li abundance as compared to single stars, and the detection of moderate Be deficiencies among Li dip stars with detectable Li abundances, all imply that that rotation induced mixing is leading to Li depletion on the main sequence. Helioseismic observations of the Sun support this hypothesis, for they show that slow form of slow mixing is operating below the base of the solar convection zone \\cite{basu}. Current stellar models which incorporate rotation induced mixing explain many, by not all of the observations. Models which are able to account for all of the data are likely to include diffusion, rotation induced mixing, angular momentum transport by gravity waves and/or magnetic fields and modest stellar winds." }, "9803/astro-ph9803330_arXiv.txt": { "abstract": "We have used a coupled time-dependent chemical and dynamical model to investigate the lifetime of the chemical legacy left in the wake of C-type shocks. We concentrate this study on the chemistry of \\HtwoO\\ and \\Otwo , two molecules which are predicted to have abundances that are significantly affected in shock-heated gas. Two models are presented: (1) a three-stage model of pre-shock, shocked, and post-shock gas; and (2) a Monte-Carlo cloud simulation where we explore the effects of stochastic shock activity on molecular gas over a cloud lifetime. For both models we separately examine the pure gas-phase chemistry as well as the chemistry including the interactions of molecules with grain surfaces. In agreement with previous studies, we find that shock velocities in excess of 10 km s$^{-1}$ are required to convert all of the oxygen not locked in CO into \\HtwoO\\ before the gas has an opportunity to cool. For pure gas-phase models the lifetime of the high water abundances, or ``\\HtwoO\\ legacy'', in the post-shock gas is $\\sim 4 - 7 \\times 10^{5}$ years, independent of the gas density. A density dependence for the lifetime of \\HtwoO\\ is found in gas-grain models as the water molecules deplete onto grains at the depletion timescale. Through the Monte Carlo cloud simulation we demonstrate that the time-average abundance of \\HtwoO\\ -- the weighted average of the amount of time gas spends in pre-shock, shock, and post-shock stages -- is a sensitive function of the frequency of shocks. Thus we predict that the abundance of \\HtwoO , and to a lesser extent \\Otwo , can be used to trace the history of shock activity in molecular gas. We use previous large-scale surveys of molecular outflows to constrain the frequency of 10 km s$^{-1}$ shocks in regions with varying star-formation properties and discuss the observations required to test these results. We discuss the post-shock lifetimes for other possible outflow tracers (e.g. SiO, \\CHthreeOH) and show that the differences between the lifetimes for various tracers can produce potentially observable chemical variations between younger and older outflows. For gas-grain models we find that the abundance of water-ice on grain surfaces can be quite large and is comparable to that observed in molecular clouds. This offers a possible alternative method to create water mantles without resorting to grain surface chemistry: gas heating and chemical modification due to a C-type shock and subsequent depletion of the gas-phase species onto grain mantles. ", "introduction": "The importance of molecular gas as the material from which stellar and planetary systems are made has led numerous investigators to examine the chemical processes that combine atoms into molecules. Over the past two decades these studies have gradually increased in both complexity and fidelity and can roughly be divided into three generations. The first generation chemical model solved the chemical rate equations at equilibrium, or steady-state, and demonstrated the importance of ion-molecule reactions in driving the gas-phase chemistry (\\cite{HK73}). The second generation of models utilized the later availability of increased computing power to study the time evolution of chemical abundances (\\cite{PH80}; \\cite{GLF82}). These models, labeled as ``pseudo-time dependent'' -- ``pseudo'' because the chemistry evolves with fixed physical conditions -- used observed physical properties of dense molecular cores ($T_k \\sim 10 - 30$ K, \\nhtwo\\ $= 10^{4-6}$ \\cc ) to show that chemical equilibrium was reached in $\\sim 10^{6-8}$ years. Pseudo-time dependent models represent the most prevalent chemical model and have been quite successful in describing the chemistry of quiescent regions in both dark and giant molecular cloud cores (\\cite{Lee_etal96}; \\cite{BGSL97}). Second generation chemical models operate under one simplifying assumption: that the molecular gas undergoes no dynamical evolution as it chemically evolves. However, molecular clouds are certainly dynamically evolving objects. The widespread occurrence of superthermal widths in molecular lines suggests that, overall, the evolution of molecular clouds is not simple quiescent evolution at a single gas temperature. Moreover, the formation of a low- or high-mass star from a molecular condensation involves a collapse that increases the density by many orders of magnitude. It has also been recognized that the birth of a protostar is associated with a period of intense mass loss, which manifests itself in energetic winds, bipolar jets, and large-scale outflows (c.f. \\cite{Lada85}; \\cite{Bachiller96}). The impact of energetic flows on surrounding quiescent gas can compress and heat the gas, in some cases providing enough energy to overcome endothermic barriers of chemical reactions, sputter or destroy grains, or even dissociate molecules. The result can be a considerably different chemical composition in the shocked gas than observed in quiescent material. Since molecules are the primary coolants of the gas, any chemical change induced by collapse or shocks can also alter the ensuing physical evolution of the cloud. To address such concerns, a third generation of chemical models has been constructed to build on the successes of previous generations by combining chemistry with dynamics. These models have been principally directed towards investigating the chemistry of core and star formation (c.f. \\cite{PTVH87}; \\cite{RHMW92}; \\cite{BL97}), the physical and chemical structure of shocked gas (c.f. \\cite{DM93}), or even complex scenarios in which gas is cycled between low and high density through recurrent episodes of low-mass star formation (\\cite{CDHW88a}). One goal of coupled astrochemical models is to search for and isolate specific molecular species that serve as signposts of particular dynamical events, such as shocks. Coupled models of shocked molecular gas have isolated one molecule in particular, \\HtwoO , which is predicted to form in large abundance in shock-heated gas (\\cite{DRD83}; \\cite{KN96a},b); if temperatures in the shocked gas exceed 400~K, the endothermic barriers of a few key chemical reactions are exceeded and all of the available oxygen not locked up in CO will be driven into \\HtwoO . This led to the assertion that water emission is a ubiquitous tracer of shock-heated gas (\\cite{NM87}). Unfortunately, due to constraints imposed by the earth's atmosphere, the detection of \\HtwoO\\ in interstellar clouds from ground-based observatories is a challenging task. Nevertheless, several studies have observed, and detected, transitions of isotopic water (\\HtwoeiO), and even \\HtwoO\\ (c.f. \\cite{Jacq_etal88}; \\cite{KL91}; \\cite{Zmuidzinas_etal94}). Quite recently, absorption from warm ($T_{gas} > 200$ K) water vapor has been unambiguously detected by the Infrared Space Observatory (ISO) towards hot stars (\\cite{Helmich_etal96}; \\cite{vDH96}), and in emission in HH54 (\\cite{Liseau_etal96}). The inferred water abundance in these sources is $\\sim 1 - 6 \\times 10^{-5}$, much larger than predicted by chemical models run for cold quiescent conditions ($x$(\\HtwoO ) $\\sim$10$^{-7}$), and is consistent with water production in warm gas either though high-temperature chemistry -- as would apply in shock-heated gas or in the near vicinity of embedded sources -- or via evaporation of water-ice mantles. An important question yet to be addressed by coupled dynamical and chemical models of shocked gas in molecular clouds is how long the high abundances of water and other molecules persist following the passage of a shock. As we will demonstrate, the time needed for shock-heated gas to return to its pre-shock temperature is several orders of magnitude less than the time required for the gas to return to its pre-shock chemical composition. Thus, the enhanced abundances of water and other species could potentially exist long after the dynamical effects of a shock passage are dissipated. Therefore, the chemistry inside a cloud is reflective of, and can be used to probe, the physical shock history of the molecular gas. Motivated by these questions, we present the results of a coupled dynamical and chemical study of molecular gas that is subjected to shocks with velocities greater than 10 km s$^{-1}$. In particular, we follow the time dependence of the chemistry in a shock-heated gas layer as it cools and the quiescent time-dependent chemistry is ultimately re-established. We present models examining this evolution using pure gas-phase chemistry as well as chemistry which includes the interaction of molecules with grain surfaces. In Section 2 we discuss the combined chemical and dynamical model. In Section 3 we use published surveys of molecular outflows to investigate the average rate at which shocks with a minimally sufficient velocity to affect the water chemistry -- 10 km s$^{-1}$ -- pass a given region of a cloud. Section 4 presents the results from our combined models along with one example of quiescent chemical evolution. Two models are presented: (1) a three-stage model of pre-shock, shocked, and post-shock gas; and, (2) a Monte-Carlo cloud simulation in which the results from Section 3 are used to examine the effects of stochastic shock activity on molecular gas over a cloud lifetime. Section 5 discusses the importance of these results on chemical models of molecular gas and also reviews the observations required to test these models. In Section 6 we summarize our results. ", "conclusions": "\\subsection{Comparison with Observations} Our results demonstrate that the {\\em time-averaged} abundance of \\HtwoO\\ is a sensitive function of the rate at which molecular gas is subjected to shocks. In order to properly interpret these results it is useful to consider how the time-averaged abundances relate to observations of water in a giant molecular cloud (GMC). The gas inside a star-forming cloud can roughly be divided into three categories: (1) quiescent material that has been relatively unaffected by current or previous epochs of star formation; (2) gas that is being physically and/or chemically affected by current star formation; and, (3) as suggested here, gas that has been affected by a previous generation of star formation and is in the process of chemically and dynamically evolving back to a quiescent state. Observations of \\HtwoO\\ toward molecular material associated with each of these categories should find abundances varying from as low as $x(\\rm{H_2O}) \\sim 10^{-6}$ in (1), and ranging upwards to $\\le 10^{-4}$ in (2) and (3). {\\em A single pointed detection of H$_2$O emission in a GMC can therefore be represented by a single stage of the three-stage or Monte-Carlo models (quiescent, shock, post-shock). Because star formation will be spread throughout a cloud or core, the time-averaged abundance therefore refers to water abundances averaged over an area that contains both quiescent material and any associated gas currently being affected by local star formation. Thus the average abundance over a cloud (cloud-average) is comparable to the time-averaged abundances in the Monte-Carlo cloud model.} This average abundance might not necessarily apply to an entire GMC complex (e.g. Orion, Gem OB1), which can extend for several square degrees on the sky, but may apply to several dense cores within a single cloud. Regions with high star-formation rates, such as OMC-1 which is associated with the Trapezium cluster (see Section 3), might be expected to have a higher probability for strong shocks, and therefore a high cloud-averaged water abundance. Cores with lower star-formation rates, such as L1641 also in Orion, would have low cloud-averaged water abundances. {\\em To test our model's ability to constrain the history of molecular clouds it is important to obtain maps of the water emission over large spatial scales in GMC cores.} Because of the strong absorption by atmospheric water vapor, observations of \\HtwoO\\ in the ISM are extremely difficult. As a result, the small number of detections of water in molecular clouds have typically been taken towards single lines of sight mostly containing luminous protostars. The very convincing detections of water by ISO provide the greatest evidence for high water abundances in hot gas, with $x(\\rm{H_2O}) \\sim 1 - 6 \\times 10^{-5}$ inferred towards hot stars (\\cite{Helmich_etal96}; \\cite{vDH96}) and HH54 (\\cite{Liseau_etal96}). Towards Sgr B2 Zmuidzinas et al. (1994) observed \\HtwoeiO\\ in absorption with the Kuiper Airbourne Observatory (KAO), while Neufeld et al. (1997) observed \\HtwoO\\ in emission using ISO. Combining these observations with previous ground based detections Neufeld et al. (1997) estimate an abundance of $x(\\rm{H_2O}) = 3.3 \\times 10^{-7}$ for the cooler outer parts of Sgr B2 and $x(\\rm{H_2O}) \\sim 5 \\times 10^{-6}$ in the hot core. These observations are important not only because \\HtwoO\\ was detected in dense gas, but also because there appears to be a range of water abundances. However, it is difficult to discern whether the enhanced water abundances are the result of high-temperature chemistry that occurs behind shocks, high-temperature chemistry appropriate to gas near young stars (c.f. Doty \\& Neufeld 1997), or evaporation of grain mantle species (see discussion in Section 5.2) and there is little information on the spatial distribution of the water emission. There have been a few attempts to map the extended emission of water. Gensheimer, Mauersberger, \\& Wilson (1996) mapped emission of the quasi-thermal $3_{13} \\rightarrow 2_{20}$ transition of \\HtwoeiO\\ in both the Orion Hot Core and Sgr B2, and found that the emission originated in a compact region ($< 10''$). Cernicharo et al. (1994) and Gonzalez-Alfonso et al. (1995) find evidence for widespread water emission in Orion and W49N using the $3_{13} \\rightarrow 2_{20}$ masing transition of \\HtwoO . In Orion they argue that the water abundance is quite high, $x(\\rm{H_2O}) > 10^{-5}$, over an extended $50'' \\times 50''$ region centered on BN-KL and the Orion Nebular Cluster. From these results, the cloud-averaged water abundance for the central regions of the Orion core near the Orion Nebular Cluster is $^{>}_{\\sim} 10^{-5}$ which, using Figures 11 and 12 (\\nhtwo\\ $= 10^6$ \\cc ; \\cite{BSG96}), is consistent with $\\tau_s \\sim 10^6$ years, as suggested in Section 3. However, the determination of abundances from masing transitions is a difficult task and these results must be viewed as suggestive until they are confirmed by mapping data obtained in other \\HtwoO\\ lines. For molecular oxygen, which also suffers from strong absorption due to atmospheric \\Otwo , the situation is even more uncertain. There exists only one tentative detection of $^{16}$O$^{18}$O towards L134N by \\cite{PLC93}, which implies $x(\\rm{O_2}) \\sim 4 - 8 \\times 10^{-5}$. However, a search for $^{16}$O$^{18}$O emission in different positions in L134N and other galactic sources by \\cite{MPLC97} did not confirm this detection and provides only upper limits of \\Otwo /CO $< 0.1$, which, assuming $x$(CO) $= 2.7 \\times 10^{-4}$ (Lacy et al 1994), gives $x(\\rm{O_2}) < 3 \\times 10^{-5}$. Searches, with similar negative results, have also been performed for molecular oxygen in extra-galactic sources with favorable redshifts (Goldsmith \\& Young 1989; Combes et al. 1991; Liszt 1992) finding $x(\\rm{O_2}) < 10^{-6}$. A recent study by Combes, Wiklind, \\& Nakai (1997) towards a z = 0.685 object provides the lowest limit to date of \\Otwo /CO $< 2 \\times 10^{-3}$. For our most realistic case, the Monte Carlo gas-grain model, we find the time-averaged \\Otwo\\ abundance is $\\sim 2 \\times 10^{-5}$. Thus, the combined effects of shocks and gas-grain chemistry could lower \\Otwo\\ abundances below the observed limits in Galactic sources, but another explanation is required to account for extra-galactic observations. The low time-averaged \\Otwo\\ abundance suggests that these results could have some bearing on the high abundances of neutral carbon relative to CO (C/CO $\\sim$ 0.1) that are observed towards a variety of star forming regions (c.f. Plume 1995, Schilke et al 1996). However, because the number of dissociative shocks which destroy CO is not high enough, the time-averaged carbon abundance in the models is well below the observed value. The best opportunity to test these results will be offered by two spaceborne observatories, {\\em The Submillimeter Wave Astronomy Satellite (SWAS)} (\\cite{Melnick_etal95}) and {\\em ODIN} (\\cite{Hjalmarson95}), both of which should launch within the next year. {\\em SWAS} and {\\em ODIN} are capable of observing and mapping the fundamental transition $1_{10} \\rightarrow 1_{01}$ of \\HtwoO\\ at 557 GHz and the $3,3 \\rightarrow 1,2$ transition of \\Otwo\\ at 487 GHz. These transitions have low upper state energies ($\\sim 26$ K) and should be readily excited in hot gas near star-forming sites and, more importantly, in the colder more extended material such as the ridge of dense gas in Orion (c.f. \\cite{Ungerechts_etal97}). \\begin{figure*} \\figurenum{16} \\plotfiddle{fig16.ps}{2.5in}{0}{50}{50}{-160}{-75} \\caption{ Points defining the plane of O$_2$ and H$_2$O abundances for the gas-grain Monte-Carlo model with $\\tau_s = 10^{6}$ years and $n_{H_2} = 10^5$ cm$^{-3}$. The spread of plotted points has been artificially increased by smearing them by 0.2 dex so as to give a better sense of the density of points in different regions of the plot. The cross represents the mean (time-averaged) abundances. } \\label{shockspec} \\end{figure*} Besides the computation of averaged abundances from mapping observations, combined observations of \\HtwoO\\ and \\Otwo\\ should be a powerful tool in examining the shock history of gas. This is demonstrated in Figure 16 where the water abundance is plotted versus the molecular oxygen abundance for the Monte-Carlo model, including grain depletion and desorption (Section 4.4.2). In this plot, the majority of points trace an area of roughly constant water abundance of $x(\\rm{H_2O}) \\sim 10^{-6}$, with the \\Otwo\\ abundance ranging between $x(\\rm{O_2}) = 3 - 30 \\times 10^{-6}$. This area has, by far, the largest number of points and defines the ``main-sequence'' of quiescent chemistry. From this main sequence a shock will trace a line extending almost horizontally to the right, with constant \\Otwo\\ abundance. Very strong shocks continue at the end of the horizontal line and extend down almost vertically, lowering the \\Otwo\\ abundance for constant \\HtwoO\\ abundance. This dependence of horizontal and vertical lines occurs because the \\HtwoO\\ is created more efficiently than \\Otwo\\ destruction (see Figures 3 and 4). This plot is a different way of examining the models, as opposed to average abundances, because it presents the evolution in a continuous fashion; the various possible solutions define a plane that traces evolutionary tracks including quiescent chemistry through a broad range of shock strengths. \\subsection{Post-Shock Chemistry and Water Ice Mantles} For comparison with previous modeling efforts, our three-stage model is similar to the episodic models presented in Charnley et al. (1988a,b). However, our model is simpler in that we do not model the formation process of a dense clump nor do we account for any mixing between shocked and non-shocked layers. Like their models, we find that the chemistry converges to well defined abundance values even when numerous cycles are repeated. Charnley et al. (1988a) also find that the water abundance varies greatly between shock and quiescent cycles, but they do not examine in detail the chemistry in the post-shock layer. An interesting result in the gas-grain models is that the abundance of water on grain surfaces in post-shock gas is quite large $x(\\rm{H_2O})_{gr} \\sim 10^{-4}$. This abundance is quite close to that inferred for water ice along lines of sight towards background stars in Taurus, where $x(\\rm{H_2O})_{gr} \\sim 8 \\times 10^{-5}$ (\\cite{Whittet93}). Thus, these results offer an alternative explanation for the large abundance of water in grain mantles -- one that requires no grain surface chemistry! We stress that this mechanism may not be responsible for all \\HtwoO\\ observed on grains. Grain surface chemistry formation of \\HtwoO\\ must be considered -- especially during cloud formation stages when the abundance of atomic hydrogen and oxygen is higher. There are also other potential desorption mechanisms that could be active and are not included in these models, such as grain mantle explosions induced by UV radiation (\\cite{SG91}) or water desorption via the infrared radiation field (\\cite{WHW92}). However, since most molecular clouds are very active star-forming sites, we suggest that caution must be applied when interpreting observations of ice mantles in molecular clouds as the sole result of grain surface chemistry. The interpretation of high gas-phase water abundances towards hot stars as the result of either high-temperature chemistry or grain mantle evaporation of \\HtwoO\\ is further obscured because water on grains may have been produced in an earlier shock episode. It is possible that the HDO/H$_2$O ratio could discriminate between water mantles created in shocks or in grain mantles and we are in the process of investigating this question (Bergin, Neufeld, \\& Melnick 1997). The large abundance of water on grains, with CO remaining in the gas phase, will also alter the ratio of total carbon-to-oxygen (C/O) in the gas phase. Comparison of chemical theory with observations of molecular abundances in GMC cores associated with massive star formation has found that the chemical abundances of many species are best reproduced with C/O ratios greater than the solar value ($>$ 0.4; \\cite{BSMP87}; \\cite{BGSL97}). For chemical modeling this is the result of depletion of the initial abundance of oxygen relative to carbon, which reduces the abundances of small oxygen-bearing species that are major carbon destroyers in the gas phase. In the three-stage gas-grain model shown in Figure 7, the C/O ratio in the post-shock gas at $t \\sim 10^{5}$ years is C/O $\\sim 0.7$. It is difficult to gauge the effect of this on other species, because of the lack of certainty with regard to high-temperature reactions. It is therefore possible that the C/O ratios inferred in Blake et al. (1987) and Bergin et al. (1997) are indicative of the core formation process, which could involve gas temperatures rising high enough to convert atomic oxygen to water. In this case, these results suggest a simple mechanism to place large amounts of oxygen on grain surfaces while still keeping most of the carbon in the gas-phase. Other molecular species, such as SO, SiO (\\cite{MPBF92}), and \\CHthreeOH\\ (\\cite{BLWC95}) have been observed with enhanced abundances in energetic outflows (see also \\cite{Bachiller96}; \\cite{Bachiller_Perez97}). These species are included in the three-stage chemical model presented in Section 4.3, but we do not present the results in detail here because of uncertainties in the high-temperature reaction rates. However, these results do have some bearing on the chemistry of these species in the post-shock layer. The formation pathway of \\CHthreeOH\\ in the gas phase is linked to a reaction of CH$_3^+$ with \\HtwoO\\ (\\cite{MHC91}). Thus, it is possible that the enhanced abundance of water through high-temperature reactions could lead to larger \\CHthreeOH\\ abundances through this reaction. In the three-stage model, the \\CHthreeOH\\ abundance in the post-shock layer is $x(\\rm{CH_3OH}) = 1 - 5 \\times 10^{-8}$ when the shock temperature ranges from 1000 to 2000 K. These abundances are at least an order of magnitude below values inferred in molecular outflows, which can be as high as $x(\\rm{CH_3OH}) \\sim 10^{-6}$ (\\cite{BLWC95}). The mechanism of methanol enhancements may therefore be the result of grain mantle evaporation or an unidentified high- or low-temperature pathway. The abundance of \\CHthreeOH\\ accreted onto grain surfaces provides another constraint. In our model the abundance of \\CHthreeOH\\ ice in the grain mantle is only 0.1 percent of the water-ice abundance. This ratio is below that observed towards NGC 7538 or W33A where $x(\\rm{CH_3OH})$/$x(\\rm{H_2O}) \\sim 10-40$ percent (\\cite{ASTH92}). Our models do not set constraints on the chemistry of SO and SiO because the large abundances of these species in outflows may be the result of sputtering of grain refractory and/or mantle material (c.f. Schilke et al. 1997; Caselli et al. 1997). When the abundances of these species are enhanced, through any mechanism, the high abundances will persist until the timescale for the individual molecule to deplete onto grain surfaces. If the dust temperature is high enough to keep a given species in the gas-phase, or for pure gas-phase chemistry, the lifetime will be $\\tau_{ra} = 4 -7 \\times 10^{5}$ years. In these models all SiO depletes onto grain surfaces at $\\tau_{dep} \\sim 10^4$ years (for \\nhtwo\\ = 10$^{5}$ \\cc\\ and assuming a sticking coefficient of unity). For \\CHthreeOH\\ abundance enhancements we find that the depletion timescale for gas-grain chemistry is equal to that of \\HtwoO\\ ($\\tau_{dep}(H_2O) \\sim 10^5$ years). The disparity in post-shock lifetimes between \\CHthreeOH , \\HtwoO , and SiO suggests that differences might exist between younger outflows, perhaps those associated with so-called Class~0 sources, and the more evolved sources (e.g. Class I). A survey of such sources in these tracers might prove to be useful in probing the links between evolutionary effects observed in outflows and the relationship to the driving source. Another possibility is that differences could also be found within different components inside a single outflow (see \\cite{Bachiller_Perez97}). The abundance of some molecular species, notably \\HCOp\\ and \\NtwoHp , are adversely affected by the high water abundance. In Figure 7, for the three-stage model, the abundance of \\HCOp\\ is decreased through reaction R6 when the water abundance is raised. Thus, the abundances of these two important molecular species should be anticorrelated. Because \\NtwoHp\\ also reacts with \\HtwoO , a similar anticorrelated behavior would be found between the abundances of \\NtwoHp\\ and \\HtwoO\\ and may account for the low \\HCOp\\ and \\NtwoHp\\ abundances in hot regions such as the Orion Hot Core (c.f. Blake et al. 1987) and in the L1157 outflow (\\cite{Bachiller_Perez97})." }, "9803/astro-ph9803324_arXiv.txt": { "abstract": "The IRAS 100 micron image of the GRB 970228 field shows that the amount of galactic dust in this direction is substantial and varies on arcminute angular scales. From an analysis of the observed surface density of galaxies in the $2.6' \\times 2.6'$ HST WFPC image of the GRB 970228 field, we find $A_V = 1.1 \\pm 0.10$. From an analysis of the observed spectra of three stars in the GRB 970228 field, we find $A_V = 1.71^{+0.20}_{-0.40}$. This value may represent the best estimate of the extinction in the direction of GRB 970228, since these three stars lie only $2.7''$, $16''$, and $42''$ away from the optical transient. If instead we combine the two results, we obtain a conservative value $A_V = 1.3 \\pm 0.2$. This value is significantly larger than the values $A_V = 0.4 - 0.8$ used in papers to date. The value of $A_V$ that we find implies that, if the extended source in the burst error circle is extragalactic and therefore lies beyond the dust in our own galaxy, its optical spectrum is very blue: its observed color $(V-I_c)_{\\rm obs} \\approx 0.65^{+0.74}_{-0.94}$ is consistent only with a starburst galaxy, an irregular galaxy at $z > 1.5$, or a spiral galaxy at $z > 2$. On the other hand, its observed color and surface brightness $\\mu_V \\approx 24.5$ arcsec$^{-2}$ are similar to those expected for the reflected light from a dust cloud in our own galaxy, if the cloud lies in front of most of the dust in this direction. ", "introduction": " ", "conclusions": "" }, "9803/astro-ph9803262_arXiv.txt": { "abstract": "We present accurate measurements of the central wavelengths of 4947 atomic absorption lines in the solar optical spectrum. The wavelengths, precise to a level $\\sim 50-150$ m s$^{-1}$, are given for both flux and disc-centre spectra, as measured in relatively recent FTS solar atlases. This catalogue modernizes existing sources based on photographic measurements and provides a benchmark to test and perform wavelength calibrations of astronomical spectra. It will also permit observers to improve the absolute wavelength calibration of solar optical spectra when lamps are not available at the telescope. ", "introduction": "\\label{sec1} Wavelength calibration is almost always needed in the process of producing useful astronomical spectra. To calibrate accurately is a non-trivial problem, in particular when working at high or very-high spectral resolution. Fourier transform spectrographs (FTS) are specially well-suited to this task, but they are not readily applied in conditions requiring high spatial or time resolution, so grating spectrometers are much more commonly used for astronomical observations. In this case, it becomes necessary to set reference positions corresponding to known wavelengths on the detector. This can be achieved by using very sharp observed telluric lines, but their location in the spectrum cannot be chosen by the astronomer. It is very usual to find spectral calibration lamps available for use with an astronomical spectrograph. The emission lines produced in the lamps have been previously measured at the laboratory, and this method usually provides a valid reference frame. However, it is often impractical to expose the calibration lamp simultaneously with the astronomical target and, unless the spectrograph is installed at a very stable focal station, the position of the spectrum on the detector varies depending on the telescope position. Accuracy is then limited by the instrument characteristics and observations of the calibration lamps are required between successive astronomical exposures. Nonetheless, calibrations via arc or hollow cathode spectra are normally accurate enough for most purposes. Ingenious techniques have been used to improve the accuracy of wavelength calibrations, such as placing gas cells at the entrance of the spectrograph (e.g., Deming \\& Plymate 1994), but it is rare to find such systems available and convenient for regular observations. On occasion the available lamps are not very rich in lines in the spectral range of observation. In some circumstances, an external check of the final precision in the translation into wavelengths would be desirable. One method for tackling problems such as these is to use solar spectra as templates. Changes in the wavelengths of the lines in the integrated sunlight spectrum around the solar cycle have been proved to be very small, bellow some 15 m s$^{-1}$ (Jim\\'enez et al. 1980; Wallace et al. 1988; McMillan et al. 1993; Deming \\& Plymate 1994). At 5000 \\AA, this translates into $\\sim$ 0.3 m\\AA, so the solar spectrum does offer a very stable source. In most practical cases, the accuracy will be imposed by the spectral resolution achieved. During night-time observations, the solar flux spectrum is observable after reflection from the Moon. Measurements of solar wavelengths in the integrated solar optical spectrum were published in 1929 by Burns and collaborators (Burns 1929; Burns \\& Kiess 1929; Burns \\& Meggers 1929), using photographic detectors and a grating spectrograph. The relatively recent solar flux FTS atlases offer a much higher quality spectrum of the Sun seen as a star. As the solar spectrum is so intense, on some solar telescopes no calibration lamps are deemed necessary, and the wavelength scale is set using the solar spectrum itself. Reasonable precision can be reached using the spectrum at the centre of the disc to compare with previously measured disc centre wavelengths, thus avoiding differential shifts due to the limb effect. In this case, small scale motions have to be averaged out, integrating in time and/or space, in order to minimize errors. The {\\it Kitt Peak Table of Photographic Solar Spectrum Wavelengths} (Pierce \\& Breckinridge 1973) has been extensively used by solar observers to set up the wavelength scale on their spectra. These observations, made on photographic plates, have been superseded in quality by the more recent FTS observations at the centre of the disc. To improve on the various sets of photographically based measurements (which date back to 1930 in the case of the solar flux spectrum), provide them in a homogeneous machine-readable format, use them to test spectral calibrations of very high resolution stellar spectra (e.g., Allende Prieto et al. 1995), and improve the accuracy of our own solar observations, we have determined the position of the central wavelengths of 4947 atomic lines in the optical solar spectrum. The employed source solar atlases, prepared from FTS data, and the fashion in which we performed the measurements is described in the succeeding sections. ", "conclusions": "" }, "9803/astro-ph9803218_arXiv.txt": { "abstract": " ", "introduction": "The goal of modern cosmology is to find the large scale matter distribution and spacetime structure of the Universe from astronomical observations, and it dates back from the early days of observational cosmology the realization that in order to achieve this aim it is essential that an accurate empirical description of galaxy clustering be derived from the systematic observations of distant galaxies. As time has passed, this realization has become a program, which in the last decade or so took a great impulse forward due to improvements in astronomical data acquisition techniques and data analysis. As a result of that an enormous amount of data about the observable universe was accumulated in the form of the now well-known {\\it redshift surveys}, and some widely accepted conclusions drawn from these data created a certain confidence in many researchers that such an accurate description of the distribution of galaxies was just about to being achieved. However, those conclusions are mainly based in a standard statistical analysis derived from a scenario provided by the standard Friedmannian cosmological models, which assume homogeneity and isotropy of the matter distribution, scenario which is still thought by many to be the best theoretical framework capable of explaining the large scale matter distribution and spacetime structure of the Universe. The view outlined above, which now has become the orthodox homogeneous universe view, has, however, never been able to fully overcome some of its objections. In particular, many researchers felt in the past, and others still feel today, that the relativistic derived idea of an eventual homogenization of the {\\it observed} matter distribution of the Universe is flawed, since, in their view, the empirical evidence collected from the systematic observation of distant cosmological sources also supports the claim that the universal distribution of matter will not eventually homogenize. Therefore, the critical voice claims that the large-scale distribution of matter in the Universe is intrinsically inhomogeneously distributed, from the smallest to the largest observed scales and, perhaps, indefinitively. Despite this, it is a historical fact that the inhomogeneous view has never been as developed as the orthodox view, and perhaps the major cause for this situation was the lack of workable models supporting this inhomogeneous claim. There has been, however, one major exception, in the form of a hierarchical cosmological model advanced by Wertz (1970, 1971), although, for reasons that will be explained below, it has unfortunately remained largely ignored so far. Nevertheless, by the mid 1980's those objections took a new vigour with the arrival of a new method for describing galaxy clustering based on ideas of a radically new geometrical perspective for the description of irregular patterns in nature: {\\it the fractal geometry}. In this review we intend to show the basic ideas behind this new approach for the galaxy clustering problem. We will not present the orthodox traditional view since it can be easily found, for instance, in Peebles (1980, 1993) and Davis (1997). Therefore, we shall concentrate ourselves in the challenging voice based on a new viewpoint about the statistical characterization of galaxy clustering, whose results go against many traditional expectations, and which keep open the possibility that the universe never becomes observationally homogeneous. The basic papers where this fractal view for the distribution of galaxies can be found are relatively recent. Most of what will be presented here is based on Pietronero (1987), Pietronero, Montuori and Sylos Labini (1997), and on the comprehensive reviews by Coleman and Pietronero (1992), and Sylos Labini, Montuori and Pietronero (1998). The plan of the paper is as follows. In section 2 we present a brief, but general, introduction to fractals, which emphasizes their empirical side and applications, but without neglecting their basic mathematical concepts. Section \\ref{Distr Galaxies} briefly presents the basic current analysis of the large scale distribution of galaxies, its difficulties and, finally, Pietronero-Wertz's single fractal (hierarchical) model that proposes an alternative point of view for describing and analysing this distribution, as well as some of the consequences of such an approach. The paper finishes with a discussion on some aspects of the current controversy about the fractal approach for describing the distribution of galaxies. ", "conclusions": "" }, "9803/astro-ph9803133_arXiv.txt": { "abstract": "\\input abstract ", "introduction": " ", "conclusions": "" }, "9803/astro-ph9803305_arXiv.txt": { "abstract": "We present an $R$-band and $J$-band photometry of an optical transient which is likely to be associated with the gamma-ray burst event GRB\\,971214. Our first measurement took place 13 hours after the gamma-ray event. The brightness decayed with a power-law exponent $\\alpha = -1.20 \\pm 0.02$, which is similar to those of GRB\\,970228 and GRB\\,970508 which had exponents of $\\alpha = -1.10 \\pm 0.04$ and $\\alpha = -1.141\\pm 0.014$ respectively. The transient decayed monotonically during the first four days following the gamma-ray event in contrast with the optical transient associated with GRB\\,970508 which increased in brightness, peaking two days after the burst, before settling to a power-law decay. ", "introduction": "The launch of the BeppoSAX satellite \\cite{boel97} in 1996 has led to a recent breakthrough in the study of gamma-ray bursts by detecting fading X-ray counterparts and localizing them to a few arc-minutes on the sky. This has allowed subsequent identification of optical counterparts in three cases, GRB\\,970228 \\cite{groo97}, GRB\\,970508 \\cite{bond97}, and GRB\\,971214 \\cite{halp97}. GRB\\,971214 triggered the BeppoSAX Gamma Ray Burst Monitor on December 14.97272 UT with a peak flux of 650 counts s$^{-1}$. In addition, the burst was localized to an error radius of \\decmin{3}{9} (99\\% confidence) with the Wide Field Camera \\cite{heis97}. The gamma-ray event was also detected by BATSE which measured a total fluence above 20 keV of $1.09 \\pm 0.07 \\times 10^{-5}$ erg cm$^{-2}$ and by one RXTE-ASM camera yielding a peak intensity of $470 \\pm 140$ mCrab \\cite{kipp97}. Shortly thereafter, a fading optical source was detected within the BeppoSAX error circle at $\\alpha(J2000)={\\rm11^h56^m}$\\decsectim{26}{4}, $\\delta(J2000)={\\rm65^\\circ12'}$\\decsec{00}{5} with I-band magnitudes of $21.2 \\pm 0.3$ on Dec 15.47 UT and $\\sim 22.6$ on Dec 16.47 UT \\cite{halp97}. Further observations by BeppoSAX detected a previously unknown fading X-ray source, later designated 1SAX J1156.4+6513, within the initial error circle at $\\alpha(J2000)={\\rm11^h56^m}25^{\\rm s}$, $\\delta(J2000)={\\rm65^\\circ13'}11''$ with an error radius of about $1'$ \\cite{anto97}. Since this second X-ray detection is consistent with the position of the fading optical source identified by Halpern \\etal\\ (1997) it is quite likely that these objects are the X-ray and optical afterglow of GRB\\,971214. We report here $R$-band and $J$-band observations of this optical transient (OT) made with the Apache Point Observatory(APO) 3.5 m telescope. This work extends the preliminary photometry reported in Diercks \\etal\\ (1997), Castander \\etal\\ (1997), and Tanvir \\etal\\ (1997). ", "conclusions": "The well-observed light curve of the GRB\\,970508 OT, the brightest observed thus far, shows a dramatic rise, peaking nearly two days after the initial burst, before beginning a power-law decay. GRB\\,970228 was not observed nearly as often through a consistent filter, but there is also evidence (depending on spectral assumptions) \\cite{guar97} that the transient increased in brightness until $\\sim20$ hr after the burst, after which it also began fading with a power-law slope essentially identical to that of GRB\\,970508. Despite the same decay slope, the GRB\\,970228 OT was $\\sim1.5$ mag fainter than the GRB\\,970508 OT. A detailed analysis of the difference between these two light curves is presented in Pedersen et al. (1998). A power-law of the form $F = F_{0}t^{\\alpha}$ was fit to the four $R$-band data points yielding $\\alpha = -1.20 \\pm 0.02$. This rate of decline is similar to the two previously identified bursts although there is no evidence of the OT brightening over the course of the observations. \\placefigure{fig-2} The observations from all four nights are combined into one deep image totaling 3.25 hrs of integration with a limiting $R$-band magnitude $\\sim25.4$ (Figure~\\ref{fig-3}). The two brightest objects within 20$''$ of the OT are an extended source (A) \\decsec{4}{6} southwest of the transient with $R = 22.7 \\pm 0.1$, and an unresolved source (B) \\decsec{4}{9} north of the transient with $R = 23.43 \\pm 0.08$. The resolution of these images is insufficient to identify any structure in the extended object. \\footnote{To obtain the images discussed in this work, contact A. Diercks or E. Deutsch}" }, "9803/astro-ph9803075_arXiv.txt": { "abstract": "We report on two observations of the Seyfert galaxy IRAS18325-5926 made in 1997 December and 1998 February with the Rossi X-ray Timing Explorer (RXTE). We find evidence for periodicities in the resulting X-ray lightcurves which are shorter than the 58~ks period found from data of the source taken in 1997 March with the imaging satellite ASCA. It is therefore likely that IRAS18325-5926 has a quasi-periodic oscillation (QPO) similar to, but at a much longer period than, the QPO seen in some Galactic Black Hole Candidates. The power spectrum of the February data has several peaks, the second highest of which is consistent with a monotonic decrease in the X-ray period. The period change is then consistent with that expected from two massive black holes spiralling together due to the emission of gravitational radiation. This possibility is very unlikely but mentioned because of its potential importance. ", "introduction": "We recently discovered a 16~hr X-ray periodicity in the Seyfert galaxy IRAS~18325-5926 using data from ASCA (Iwasawa et al 1998). Nearly 9 cycles of the oscillation were observed, with a total amplitude of about 15 per cent. The active nucleus has similar properties to that of Seyfert 1 galaxies, except that the power-law continuum is slightly steeper than most (photon index $\\Gamma\\sim 2.1$) and there is moderate absorption by a column density of $\\sim 10^{22}\\pcmsq$. There is also a broad iron line in the X-ray spectrum (Iwasawa et al 1996) which indicates the presence of an accretion disk in the X-ray emission region, viewed at moderate inclination (40--50 deg). The periodicity is plausible for Keplerian motion at 10--20 gravitational radii around a black hole of mass $2\\times 10^8 - 2\\times 10^7\\Msun$. It also scales well with quasi-periodic oscillations (QPO) seen from some Galactic Black Hole Candidates (BHC; Belloni et al 1997). The cause of the oscillation in IRAS~18325-5926, or indeed in any BHC, is unknown. In order to determine whether the variation is exactly periodic or a QPO, we observed IRAS~18325-5926 with the Rossi X-ray Timing Explorer in 1997 December 25--27 and 1998 February 21--23. We report here the light curves and power spectra of those observations, which also show oscillations, but of different periods. We conclude that the AGN has a clear QPO signal and note that our results are consistent with the exciting possibility that 2 black holes are rapidly spiralling together. ", "conclusions": "\\begin{table} \\begin{center} \\caption{Mean flux and period for X-ray observations of IRAS~18325-5926. The Ginga data are reported by Iwasawa et al (1995). The error bars indicate the period change over which the power drops by a factor of two from the peak. 1998a and b refer to the highest and next highest peaks in the February power spectrum, respectively.} \\begin{tabular}{ccr} Detector & Flux (4--10~keV) & Period \\\\ & $\\ergpcmsqps$ & ks \\\\[5pt] Ginga 1989 & $1.5\\times 10^{-11}$ & $>30$ \\\\ ASCA\\ 1994 & $5.6\\times 10^{-12}$ & $>80$ \\\\ ASCA\\ 1997 & $1.0\\times 10^{-11}$ & $58.0^{+2.6}_{-1.8}$ \\\\ RXTE\\ 1997 & $2.6\\times10^{-11}$ & $38.7^{+4.2}_{-2.9}$ \\\\ RXTE\\ 1998a & $2.1\\times 10^{-11}$ & $40.3^{+4.5}_{-2.7}$ \\\\ RXTE\\ 1998b & $2.1\\times 10^{-11}$ & $28.2^{+1.9}_{-1.4}$ \\\\ \\end{tabular} \\end{center} \\end{table} The X-ray emission from IRAS~18325-5926 oscillates in a manner similar to that seen in BHC QPO. Such a large clear oscillation from matter around a black hole may suggest that we are detecting the fundamental frequency of some space-filling corona above the disk. This may be possible if the corona has a proton temperature close to the virial value and a lower electron temperature (see e.g. Di Matteo, Blackman \\& Fabian 1997). Of course this must happen over a very restricted range of radii in order that a single dominant oscillation is seen. Perhaps it corresponds to the radius where the surface emission from the disk peaks (i.e. $\\sim$7 Schwarzschild radii for a non-spinning hole). The variation of period with flux (Fig.~5) is similar to that seen in some QPO. Otherwise, as mentioned by Iwasawa et al (1998), it could be due to the Bardeen-Petterson effect if the angular momentum vectors of the black hole and accreting matter are not aligned. A range of radii are then selected over which the accretion disk tilts over to match the equatorial plane of the black hole. (The disk is not actually precessing in this case; Markovic \\& Lamb 1998.) Reflection/obscuration by blobs of gas in the tilt zone might then lead to observed flux variations, especially if the inclination is fairly high (as the broad iron line may indicate; Iwasawa et al 1996), but we consider it unlikely that 50 per cent variations can be obtained in this way. \\begin{figure} \\centerline{\\psfig{figure=pflux.ps,width=0.48\\textwidth,angle=270}} \\caption{Variation of period with 4--10~keV flux. The two highest power spectrum peaks are shown for February.} \\end{figure} The data are consistent with a continuous decrease of period with time from GINGA observations in 1989 to December 1997, and continuing if the second highest peak in the February power spectrum is used. This raises the possibility that the period is due to a massive object orbiting the black hole (see Cunningham \\& Bardeen 1973). The separation of the objects requires that this second object is compact. The period decrease is then explained as due to orbital energy loss by the emission of gravitational waves (as for the binary pulsar). The rate of period change $$\\dot P ={{-96}\\over5}{{G^{5/3}}\\over c^5}{{M^{2/3}\\mu(4\\pi^2)^{4/3}}\\over P^{5/3}},$$ where $M=M_1+M_2$, the sum of the separate masses, and $\\mu=M_1M_2/(M_1+M_2).$ This integrates to give $P\\propto(t_0-t)^{3/8}$ where $t_0$ is the time when the masses finally merge. We have fitted this last relation to the data and find an acceptable fit (Fig.~6). It suggests that the merger may occur in late April 1998. The rate of spiral-in enables us to estimate that the product $M^{2/3}\\mu\\approx 1.5\\times 10^{10}$ where the masses are in units of $\\Msun$. This means that for $M_1\\sim 2\\times 10^6\\Msun - 10^8\\Msun$, $M_2>10^5\\Msun$. There is no solution for $M_1<1.5\\times 10^6\\Msun$. Much of the energy from such a merger would emerge as gravitational waves, but some is likely throughout the electromagnetic spectrum. We note that it is {\\it a priori} unlikely to find an object close in time to a spiral-in merger event. Possibilities that can enhance that probability are if a) black holes are built out of many merger events where the basic unit has a mass of only a few $10^5\\Msun$ (say from dwarf galaxies) and b) the inspiralling black hole switches on, or significantly enhances, an otherwise quiescent active nucleus. If we make the extreme argument that this process happens in all galaxies of space density $3\\times10^{-3}n_{-2.5}\\Mpc^{-3}$, all of which have a central black hole of mass $5\\times10^7M_{7.5}\\Msun$ growing by the addition of smaller black holes of $10^5m_5\\Msun$, then a merger will take place within $120\\Mpc$ (the distance to IRAS18325-5926) every $1000m_5 M_{7.5}^{-1} n_{-2.5}^{-1}\\yr$. We therefore see that such an event is not completely improbable but is unlikely. The probability that we should find the signal of a merger in its last year (it was the ASCA result of an observation in 1997 March which alerted us to make the RXTE observations) is $\\sim 10^{-3}$. Such mergers would of course be more common at fainter flux levels ($\\sim 1\\yr^{-1}$ within 1 Gpc), and the prospects for space-based gravitational wave astronomy (which will be sensitive to events in massive black holes) could be very positive. A lower mass black hole captured by a central black hole is likely initially to have a highly eccentric orbit (Sigurdsson 1997). When the eccentricity $e$ is high, the above estimates for $\\dot P$ are dramatically increased (Shapiro \\& Teukolsky 1983), for example by a factor of 100--1000 when $e\\sim 0.8-0.9$, respectively. This allows $M_2$ to be much lower than estimated above for a circular orbit (it can be as low as about $100\\Msun$). The overall temporal behaviour of the system is then more complex ($e$ decreases with time). It may be difficult for a low mass black hole to modulate the observed X-ray emission greatly; this problem is offset by the enhanced probability of such an event occurring. We stress that this last, exciting, interpretation involving an in-spiralling black hole is unlikely and depends upon the identification of the second highest power peak in the February 1998 data with the orbital period of the second black hole. It does not explain the other peaks in the power spectrum nor why some peaks expected in the lightcurve do not occur. We discuss it here only because of its potential importance. Clearly further observations are urgently required. Even if such observations fail to show a consistently decreasing period, IRAS1832-5926 has a convincing, high-amplitude, QPO, which requires explanation. \\begin{figure} \\centerline{\\psfig{figure=spiral_in_new.ps,width=0.48\\textwidth,angle=270}} \\caption{The peak period displayed against time in days, measured backwards from the February 1998 observation. Error bars show where the peak power has decreased by about a factor of 2. The line shows the best-fitting relation $P=6203.(t+63.5)^{3/8}\\s$. For the February data we plot both the period with the peak power and the longer one which had slightly less power.} \\end{figure}" }, "9803/astro-ph9803243_arXiv.txt": { "abstract": "We report the discovery of two low redshift HI 21cm absorbers, one at $z = 0.2212$ towards the $z_{em} = 0.630$ quasar OI 363 (B0738+313), and the other at $z = 0.3127$ towards PKS B1127-145 ($z_{em} = 1.187$). Both were found during a survey of MgII selected systems at redshifts $0.2 < z < 1$ using the new UHF-high system at the Westerbork Synthesis Radio Telescope (WSRT). New HST/FOS observations also identify both systems as damped Ly$\\alpha$ (DLa) absorbers. By comparing the column density from the DLa line with that from the HI 21cm line, we calculate the spin temperature, T$_s$ of the two systems. We find $T_s \\approx 1000$ K for both of these low redshift absorbers. For the $z = 0.3127$ system towards PKS B1127-145, two galaxies have been previously identified with emission lines at the absorber redshift (Bergeron \\& Boiss\\'e, 1991), with the galaxy at a closer projected distance to the quasar assumed to be responsible for the absorption system. An ESO-NTT/EFOSC2 spectrum of a 3rd, fainter companion at 3.9 arcsec or 11 $h^{-1}_{100}kpc$ from the line of sight of PKS 1127-145 reveals [OIII]4958 and 5007 at $z = 0.3121 \\pm 0.0003$. We consider this object the most likely to be responsible for the 21cm absorption, as it is much closer to the QSO sightline than the two galaxies identified by Bergeron \\& Boiss\\'e. \\end {abstract} ", "introduction": "The study of low redshift examples of the quasar absorption line systems responsible for the damped Ly$\\alpha$ (DLa) and HI 21cm absorption lines is important to help bridge our understanding of neutral gas-rich systems between those at redshift $z = 0$ and those at high $z$. Our knowledge of the neutral gas in nearby spiral galaxies is mainly based on observations of the HI 21cm line in emission. At high redshift, however, we observe the HI 21cm line in absorption, which can only be seen along a limited number of lines of sight through the intervening absorber, making detailed knowledge of the gas characteristics difficult. The low redshift ($z < 1$) neutral absorbers are still close enough that both optical and radio data of reasonable quality can be obtained in order to understand their kinematics and physical gas characteristics. Such information is necessary to build a framework for a correct interpretation of the higher z counterparts to these systems. Searches for redshifted HI 21cm absorption can be time-consuming since radio spectrometers typically observe only relatively narrow instantaneous bandwidths and only the highest column density QSO absorption line systems have measurable optical depths in the 21cm line. Since DLa systems have high column densities of neutral HI, they are the most likely objects to have HI 21cm absorption. Unfortunately for low redshift work, the Ly$\\alpha$ line is not shifted into the optical window until $z \\simeq 1.65$, so finding these lines requires UV spectra to be taken with space telescopes. This, combined with the small cross section for DLa absorption, means that only a small number of DLa systems have been identified at low redshift. Therefore, alternative selection criteria which are reasonably effective at finding an HI 21cm absorber must be used. All known DLa and HI 21cm absorbers have associated low-ionization metal lines (cf. Lu and Wolfe, 1994) such as the MgII $\\lambda\\lambda$2796, 2803 doublet, which can be observed easily in ground-based spectra down to about $z = 0.1$. A study of MgII selected systems using previously existing UV data yielded about 1 DLa system per 10 MgII systems observed (Rao, et al., 1995). A similar statistic exists for HI 21cm absorption in MgII systems (Briggs \\& Wolfe, 1983). This suggests that MgII can be used to optically select systems likely to have high column densities of neutral gas observed either as DLa or HI 21cm absorption. We are conducting a survey for HI 21cm absorption in low redshift MgII selected systems using the Westerbork Synthesis Radio Telescope (WSRT). In this paper we present two new HI 21cm absorbers from our survey, one towards PKS B1127-145 at $z = 0.3127$ and the other towards OI 363 at $z = 0.2212$. Recent HST/FOS spectra have identified DLa absorption in both of these objects as well. ", "conclusions": "\\subsection{OI 363} OI 363 (0738+313), is a core dominated quasar at $z_{em} = 0.630$. Observations at 1640 MHz (Murphy, et al., 1993) show that the lobes extending $\\approx 30\\arcsec$ from the core contain only $3\\%$ of the total flux of the quasar. The quasar is slightly variable at low frequencies (Bondi, et al., 1996b). The metal line absorption system was originally reported by Boulade, et al. (1987) at a redshift of $z = 0.2213$, and subsequently by Boiss\\'e, et al. (1992) at a redshift of $z = 0.2216 \\pm 0.0003$. The only identified lines in the spectrum are the MgII $\\lambda\\lambda2796,2803$ doublet and a possible MgI line. Deep optical imaging was made by LeBrun et al. (1993), who identified what they considered a likely absorber at a projected separation from the quasar of 5.70'', or $1.24R_H$ if it lies at $z = 0.221$. They identified three additional galaxies, at smaller angular separations from the quasar, which are very faint and would be dwarf galaxies at the absorption redshift. Unfortunately, they do not report a confirmed redshift for any object in the field near the quasar. The HI 21cm absorption line, shown in fig. 1, has a narrow width of only two channels over most of its depth, and may be unresolved by this observation. This implies an upper limit to the line width of $\\sim8$ km s$^{-1}$ at the 6 km s$^{-1}$ resolution of the data. The HI column density from the DLa profile is $N(HI) = 7.9\\pm1.4 \\times 10^{20}$ cm$^{-2}$, and the calculated mean harmonic spin temperature is T$_s = 1230 \\pm 335$ K. The termal kinetic temperature of the gas for a line of width 8 km s$^{-1}$ is T$_k = 1400$ K. This means that within the errors, T$_s = $T$_k$. \\subsection{PKS B1127-145} PKS B1127-145 is a compact, gigahertz peaked radio source at $z = 1.187$. VLBI observations at 1670 MHz show a slightly elongated structure with an extent of approximately 20 mas. Observations at 408 MHz give a flux variability of 1.2 Jy/year (Bondi, et al., 1996a), and indicate structural variations as well. Fig. 2 shows the neutral HI 21cm absorption for this system. The observed flux of the quasar is 5.25 Jy in this observation. This is somewhat lower than reported fluxes in NED\\footnote{The NASA/IPAC Extragalactic Database (NED) is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration.}, however, as discussed above the quasar is a known variable at low frequency. If the absorption line is fit by one gaussian component, the optical depth of the absorption is $6.2\\%$, and the FWHM is approximately 42 km s$^{-1}$. However, there is some evidence for structure in the line. In particular, the split in the middle appears to be real, suggesting at least two components, and the asymmetric low frequency side of the profile may result from a third component. There is some low level interference in the spectrum adjacent to the low frequency side of the absorption line, at $\\sim$1081.75 MHz, which makes interpretation of the real shape difficult. The HI column density from the DLa profile is $N(HI) = 5.1\\pm0.9 \\times 10^{21}$ cm$^{-2}$, and the calculated T$_s = 1000 \\pm 200$ K. Bergeron and Boiss\\'e (1991) studied this $z_{abs}=0.313$ metal-line system in some detail. They identify MgII, FeII and MgI absorption in the spectrum. The average redshift of the metal lines is $z = 0.3127 \\pm 0.0002$. In addition to the absorption spectrum, and deep images of the field, they present emission spectra for two of the bigger bright galaxies near the quasar (different from the two faint companions discussed in Sect. 4), both of which have redshifts similar to that of the metal line system. They identify the galaxy at the smaller projected distance from the quasar sightline as the absorber. The ESO/EFOSC2 spectral observations (discussed in Sect. 3) identify a new candidate absorber galaxy for this system. The new emission object is a close companion to the west of the quasar and lies at a projected distance of $3.9\\arcsec$ or 11 $h^{-1}_{100}$kpc from the quasar sightline. Based on comparison with other objects in our image for which apparent magnitudes are given in Bergeron and Boiss\\'e (1991), it has an apparent magnitude of $m_r = 22.3$. The galaxy which Bergeron and Boiss\\'e identify as the absorber is at a projected separation of $9.6\\arcsec$, or 37 $h^{-1}_{100}$kpc from the quasar at the redshift of the galaxy. This corresponds to a galactic radius of $2.7 R_H$. A column density of neutral gas of $10^{21}$ cm$^{-2}$ is unlikely at this galactic radius, and we consider the new emission object, although smaller and fainter, to be the more likely absorber given its proximity to the quasar sightline. It is also possible that the three galaxies at similar redshift have undergone strong interaction with each other, in which case the absorbing gas could be tidal debris. \\subsection{Spin Temperature} Fig. 4 shows T$_s$ vs. redshift for all of the known HI 21cm and DLa absorber systems, calculated from the literature (as summarized by Carilli, et al., 1996), with the two new data points marked by open symbols. The shaded region shows the range of Galactic T$_s$ values at optical depths comparable to those found in the DLa systems, using numbers from Braun and Walterbos (1992). The large errorbars in any given measurement are dominated by uncertainties in the true optical continuum level, due to confusion from the Ly$\\alpha$ forest lines, which make fitting the damped profile difficult. As noted in previous studies (cf. de Bruyn, et al. 1996), all of the redshifted absorbers except the highest optical depth (lowest T$_s$) system have T$_s$ values which are roughly two or more times greater than the Galactic values at similar optical depths. Most estimates of T$_s$ for the present epoch have been based on studies of the Milky Way or Andromeda (cf. Dickey and Brinks (1988); Braun and Walterbos (1992), and references therein). These estimates use column densities derived from HI 21cm emission lines rather than from DLa absorption lines, but studies have shown that N(HI)$_{Ly\\alpha} \\approx$ N(HI)$_{21cm}$ (Dickey and Lockman, 1990) . This suggests that T$_s$ values for the galaxy and the redshifted DLa systems, although calculated from different quantities, can be compared. On the other hand, values for T$_s$ have a strong correlation with the column density, N(HI), in each of the clouds along the line of sight in the Galaxy, and are thus sensitive to the location of individual gas clouds. A single cloud usually has a greater angular size on the sky than the beam with which it is observed, so it is possible to calculate the spin temperature for just one cloud. This is not the case in redshifted systems, where a line of sight includes more of the galaxy and possibly many clouds. Thus it is difficult to usefully compare present epoch spin temperature values to those calculated at higher redshifts. The new systems in this study were observed at low redshift with the same line of sight limitations as the higher redshift systems, allowing a meaningful comparison. There is no clear trend for a change in T$_s$ with increasing redshift among the eight systems shown in Fig. 4. If the gap between T$_s$ values in the galaxy and those in the DLa systems is an effect of evolution over time, then all of that evolution must have occured between $z = 0.2$ and the present epoch. Instead, we consider it more likely that the gap arises from the presence of many clouds in one line of sight at high redshift, or from differences in the radio and optical sightlines. The sizes of the radio emission regions of quasars are much larger than the optical regions, and usually larger than an average cloud as well. It is therefore likely that the optical and radio lines of sight actually sense different clouds in a redshifted galaxy, and hence have different column densities of neutral gas. This implies that the spin temperatures derived by assuming both column densities are equal may be meaningless. Unfortunately, the resolution obtained in most radio survey observations (with single dishes or synthesized beams from arrays like the WSRT) is at least as large on the sky as compact background radio sources, and gives no spatial information about the clouds in front of the quasar. Future use of VLBI techniques to pinpoint the radio sightlines more accurately may help to clarify this problem." }, "9803/hep-ph9803295_arXiv.txt": { "abstract": "The effects of a possible rotation of the galactic dark halo on the calculation of the direct detection rates for particle dark matter are analyzed, with special attention to the extraction of the upper limits on the WIMP--nucleon scalar cross section from the experimental data. We employ a model of dark halo rotation which describes the maximal possible effects. For WIMP masses above 50 GeV, the upper limit exclusion plot is modified by less than a factor of two when rotation is included. For lighter masses the effect can be stronger, suggesting the necessity to develop specific models of halo rotation in order to provide more accurate conclusions. ", "introduction": "The possibility to detect Weakly Interacting Massive Particles (WIMPs) distributed in the halo of our Galaxy has been a major issue in the last years, since these particles could provide the amount of dark matter necessary to explain many observed dynamical properties of galaxies, clusters and of the Universe itself. Different kinds of possible signals have been identified and looked for, in order to outline the presence of WIMPs in our Galaxy. These signals are usually referred to as ``direct\" and ``indirect\" detection rates. Direct detection refers to the possibility to measure a WIMP--nucleus interaction in a low--background detector, while indirect detection relies on the measurement of WIMP annihilation products: photons, antiprotons and positrons produced by the annihilation in the galactic halo, or neutrinos coming out of the Earth or the Sun where WIMPs may have been accumulated as a consequence of gravitational capture. It is remarkable that the present sensitivity of the different experiments is already at the level of the predicted rates for specific WIMP candidates, like the neutralino, which represents one of the most interesting and studied cold relic particles \\cite{noi}. The calculation of the different detection rates depends not only on the particle physics properties of the WIMPs interactions, but also on the characteristics of the galactic halo where the WIMPs are distributed. Direct detection rates and upgoing-muon fluxes at neutrino telescopes, which both rely on the WIMP elastic scattering off nuclei, depend on the WIMP matter density $\\rho_\\odot$ and velocity distribution $f_\\odot(v)$ at the Earth position $r_\\odot$ in the Galaxy. In particular, the dependence of the signals on $\\rho_\\odot$ is a linear one. The other indirect signals (photon, antiproton and positron fluxes) have a stronger dependence on the matter distribution, since they are proportional to the square of the matter distribution function (DF) $\\rho(\\vec r)$ integrated over the effective region of production and propagation of the annihilation products. On the contrary, this kind of signals are essentially independent on the details of the velocity DF, since the annihilating WIMPs are almost at rest and corrections due to their velocity dispersion are negligible. Detailed estimates of the detection rates would require specific and accurate models of the galactic halo able to provide a reliable WIMPs DF $g(\\vec x, \\vec v)$ (not necessarily separable in phase space, i.e. $g(\\vec r, \\vec v) = \\rho(\\vec r) f(\\vec v)$). Unfortunately, detailed halo models are not available at present, mainly because the constraints obtained from astrophysical observations are not stringent enough to restrict different possibilities. The most important observational constraint is provided by the flatness of the rotation curves at large radii. Although the available data on our Galaxy do not provide a compelling evidence of a flat rotation curve, this feature is observed in a large number of spiral galaxies and therefore it looks reasonable to assume its validity also for our Galaxy. The standard and simplest model of the dark galactic halo, which is compatible with a flat rotation curve, is the so--called isothermal sphere. This model relies on the two basic assumptions of spherical symmetry and thermal equilibrium, which find a strong support in the argument of ``violent relaxation\" introduced by Lynden--Bell 30 years ago\\cite{lynden-bell}. In this model, the DF is separable into a matter density distribution $\\rho(r)$, which has a $r^{-2}$ behaviour at large radii, and into a Maxwell--Boltzmann (MB) velocity DF $f(v)$ \\cite{binney}. Although such a model gives a divergent total mass and therefore an appropriate cut-off has to be introduced at large radii, its range of validity has been tested at least in the inner parts of many galactic systems. Moreover, since it represents a simple and reasonable approximation, in the absence of a more detailed model it is widely adopted to describe the dark halo of our Galaxy. However, many different models are known to be consistent with flat rotational curves. For instance, models which describe non--spherically symmetric or flattened halo distributions have been discussed \\cite{binney}. In these models, the specific form of $\\rho(\\vec r)$ differs from the standard isothermal sphere matter DF, especially at small radii, entailing quite large uncertainties on the local value $\\rho_\\odot$. A comprehensive numerical study which takes into account a large number of models indicates that the local value of the non--baryonic dark matter density falls in the (rather conservative) range 0.1 $\\lsim \\rho_\\odot \\lsim$ 0.7 GeV cm$^{-3}$\\cite{turner}. Contrary to the matter DF, the specific form of the velocity DF $f(v)$ has been much less investigated. Modifications to the standard MB velocity DF are known \\cite{binney,evans}, but the problem of determining the correct form of the distribution of the WIMP velocities in the halo has no clear and simple solution at present, both theoretically and observationally. The velocity DF is required to be consistent with a given $\\rho(\\vec r)$ but this, in general, does not determine $f(v)$ in a unique way. The calculation of the WIMP detection rates is usually performed by using the standard isothermal sphere model. However, modifications in the isothermal model can affect the detection rates, introducing uncertainties in the theoretical predictions and in the extraction of the experimental limits on the WIMPs parameters. The effects induced on the detection rates by a modification in the matter DF are simple to take into account, since the dependence of the detection rates on $\\rho(\\vec r)$ can be factorized. Specifically, the physical range of $\\rho_\\odot$ quoted above implies an uncertainty of about a factor of 7 in the evaluation of the direct detection rates and in the neutrino fluxes \\cite{noi} (it has to be remarked that this large factor reflects a rather conservative attitude). Even larger uncertainties affect the indirect rates from WIMP annihilation in the halo, since in this case a modification in the matter density profile can strongly affect the integral of $\\rho^2(\\vec r)$ over the effective production region of the signal \\cite{pbar_to,pbar_japan,gamma}. Contrary to the case of the matter DF, a modification of the standard MB velocity DF would affect the direct detection rates and the indirect rates at neutrino telescopes in a much more involved way. This is because the dependence of these rates on $f(v)$ is through a convolution of $f(v)$ with the differential WIMP--nucleus cross section. Since the WIMP--nucleus scattering depends on the relative velocity of the WIMPs with respect to the detector nuclei, a potentially significant effect could be due to a bulk rotation of the halo. This would necessarily modify the WIMP phase--space DF with respect to the standard MB form. In this paper we wish to discuss the possible effects induced by a halo rotation on the direct detection rates, with special attention to the ensuing consequences on the determination of the upper limits on the WIMP--nucleus cross section from the experimental data. A calculation of the direct detection rates in the case of a rotating halo has been addressed in Ref. \\cite{kamion}, where it has been concluded that the maximal effect of rotation leads to a 30\\% effect on the total detection rates for a Ge nucleus, in the case of an ideal detector with no threshold. However, when considering a real detector the behaviour of the differential rates at threshold and the detector characteristics are crucial in determining the experimental limits on the WIMP--nucleus cross section \\cite{noi}. Therefore, we explicitly take into account the features of running detectors, such as thresholds, quenching factors and energy resolution, in order to estimate the largest uncertainties induced by a possible halo rotation in a confident way. To this aim, following Ref.\\cite{kamion} we model the galactic rotation as described by Lynden--Bell in Ref.\\cite{lynden-bell2}, where the maximally rotating velocity DF compatible with a given mass distribution has been derived, on the ground of purely kinematical arguments. Even if Lynden--Bell's model of halo rotation may not represent a situation which is realized in a physical halo, we consider it useful to bracket the size of the effect of halo rotation on the direct detection rates. The plan of our paper is the following. In Sect.II we briefly describe the calculation of the direct detection rates in the presence of halo rotation. In Sect.III we discuss our results for Ge, NaI and Xe detectors, taking into account the most recent experimental data of the different Collaborations. Finally, in Sect.IV we draw our conclusions. An Appendix is added, where we report the analytical expressions of the relevant part of the direct detection rates which contain the details of the velocity DF in the case of the standard non--rotating, maximally co--rotating and maximally counter--rotating haloes. ", "conclusions": "In this paper we have investigated the effect induced by a possible rotation of the galactic halo on the rates of WIMP direct detection. In particular, we have discussed the implication of halo rotation on the determination of the exclusion plots on the WIMP--nucleon cross section for different detectors, namely Ge, NaI and Xe ones. The rotation of the halo has been described by using a model \\cite{lynden-bell2} which corresponds to a situation where the halo possesses the maximal rotation compatible with a given mass DF, which for simplicity we have chosen to be that of the isothermal sphere. We found that the exclusion plots obtained from the data are affected by less than a factor of 2 in the case of counter--rotating models. The same size of uncertainty occurs also for the co--rotating models, when the WIMP mass is larger than about 50 GeV. For lighter WIMPs and co--rotation, the exclusion plots are modified by a larger amount. We have to remind that, due to the particular model of halo rotation which we have employed here, these are expected to be maximal effects. For specific physical rotation models, the effect of halo rotation will be plausibly smaller. We can therefore conclude that, at least for WIMP masses greater than about 50 GeV, the determination of the exclusion plots from the experimental data are affected by an uncertainty smaller than a factor of two due to the possibility that the galactic halo rotates, independently on the specific model of halo rotation. We notice that recent preliminary data from accelerators indicate that the lower limit on the mass of the most plausible WIMP candidate, the neutralino, is $m_\\chi\\simeq 30$ GeV for low value of the susy parameter $\\tan\\beta$, and $m_\\chi\\simeq 45$ GeV for $\\tan\\beta\\gsim 3$ \\cite{lep183}. Therefore, for this dark matter candidate, the uncertainty on the exclusion plot due to a possible rotation of the halo is expected to be relatively small. The situation is different for lighter WIMPs. In this case, it would be required to develop specific models of halo rotation in order to obtain more accurate conclusions." }, "9803/astro-ph9803239_arXiv.txt": { "abstract": "We report the discovery of a 2.1hr optical modulation in the transient source GS1826-24, based on two independent high time-resolution photometric observing runs. There is additional irregular variability on shorter timescales. The source also exhibited an optical burst during each observation, with peak fluxes consistent with those of the three X-ray bursts so far detected by {\\it Beppo}SAX. We compare the low-amplitude variation ($\\sim 0.06^m$) to that seen on the orbital periods of the short period X-ray bursters, X1636-536 and X1735-444, as well as the similarity in their non-periodic fluctuations. Other transient neutron star LMXBs possess short periods in the range 3.8-7.1 hrs. However, if confirmed as the orbital, a 2.1 hr modulation would make GS1826-24 unique and therefore of great interest within the context of their formation and evolution. ", "introduction": "GS 1826-24 was discovered serendipitously in 1988 by the {\\it Ginga} LAC during a satellite manoeuvre (Makino {\\it et al.} 1988). The source had an average flux level of 26 mCrab (1-40 keV), and a power law spectrum with $\\alpha = 1.7$. Observations both a month before and after by the {\\it Ginga} ASM, by TTM in 1989 \\cite{intZ92} and by ROSAT in 1990 and 1992 \\cite{barr95} found comparable flux levels. Temporal analyses of both the {\\it Ginga} detection and ROSAT data yielded a featureless $f^{-1}$ power spectrum extending from $10^{-4} - 500$ Hz \\cite{tan95,barr95}, with neither QPO nor pulses being detected. Despite its detection by {\\it Ginga}, the source had not been previously catalogued. Neither were X-ray bursts detected by {\\it Ginga}. Together with its similarities to Cyg X-1 (hard X-ray spectrum, strong flickering), this led to an early suggestion by \\scite{tana89} that it was a soft X-ray transient with a possible black-hole primary. Later, \\scite{stri96} called this suggestion into doubt, following examination of data from CGRO/OSSE observations. They found that fitting both the Ginga and OSSE spectra produced a model with an exponentially cutoff power law plus reflection term. The observed cut-off energy of $\\sim$58 keV is typical of the cooler neutron star hard X-ray spectra. The recent report of three X-ray bursts detected by {\\it Beppo}SAX \\cite{uber97} and our detection of optical bursts here confirms the presence of a neutron star accretor. Following the first ROSAT/PSPC all-sky survey observations in September 1990, and the determination of a preliminary X-ray position, a search for the counterpart yielded a time variable, UV-excess, emission line star \\cite{motc94,barr95}. This source had $B= 19.7 \\pm 0.1$, and an uncertain V magnitude of $V \\simeq 19.3$, due to contamination by a nearby star. There was also evidence for $\\simeq 0.3^m$ variations on a one hour timescale, but the time sampling was fairly poor. For this reason, we included this object in our target list for a high-speed photometry run at the South African Astronomical Observatory (SAAO). Further time-series photometry was also obtained on the William Herschel Telescope, La Palma (WHT) to confirm the variability that was seen. ", "conclusions": "\\begin{table*} \\begin{minipage}{150mm} \\caption{Properties of GS1826-24 and the short-period transient bursters$^a$\\label{tab:sptb}} \\footnotetext[1]{Data taken from \\scite{vP95} and references therein, unless cited elsewhere.} \\begin{tabular}{c c c l c l l l } \\hline Source \t& $P_orb$ & V \t\t& B-V, U-B \t&\tE${\\rm _{B-V}}$\t\t& F$_X$\t&\tActive &Quiescent \t\t\t\\\\ &\t(hr)\t&\t\t\t&\t& & ($\\mu$Jy)&\t& \\\\ \\hline GS1826-24&\t2\\footnote[2]{This work.}\t\t&19.3\t\\footnote[3]{\\scite{motc94}.}\t& 0.4$^c$, -0.5$^c$\t&\t0.4$^c$\t& 30$^c$& 1988-present\\footnote[4]{\\scite{mak88}.} & before 1988\\\\ &\t\t\t&\t\t\t&\t&\t\t\t&\t\t&&\\\\ X0748-678&\t3.82\t&$>$23\\footnote[5]{\\scite{wad85}.}-16.9&0.1, -0.9\t&0.42\t\t& 0.1-60\t& 1985-present\\footnote[6]{\\scite{whi95}.} &before 1985\\\\ &\t\t\t&\t\t\t&\t\t&\t\t&\t&\t\\\\ X2129+470 &\t5.24\t&16.4-17.5&\t0.65,-0.3 &0.5\t\t&\t9\t& $<$1979 - 1983$^f$ & 1983-present\\\\ &\t\t\t&\t\t\t&\t&\t\t&\t\t\t&\t\t&\\\\ X1658-298 & 7.11& 18.3\t\t&0.45,-0.4\t\t& 0.3 & $<$5-80& pre-1980s $^f$ & 1980-present\t\\\\ &\t\t\t&\t&\t&\t\t\t&\t\t&&\\\\ \\end{tabular} \\end{minipage} \\end{table*} The period distribution of low-mass X-ray binaries (LMXB) has shown a scarcity of systems below 3 hr presumably in part due to their faintness, but there is a notable absence of any systems with periods between \\til 1 and 3 hr \\cite{whi85a,whi85b}. Recently, \\scite{king97} have investigated the formation of neutron star LMXBs. They found that in order to produce the relatively large fraction of soft X-ray transients in the $\\simlt$ 1-2 day range, the secondaries must have 1.3\\msun $\\simlt M_2\\simlt$ 1.5\\msun\\ at the onset of mass transfer and be significantly nuclear evolved (provided that the SN kick-velocity is small compared to the pre-SN orbital velocity). This mass range ensures that the mass transfer rates driven by angular momentum loss are below the critical rate needed for SXT behaviour in the required fraction of neutron star LMXBs. The large initial masses then account for the rarity of any short $P \\simlt 3$ hr systems. If indeed the 2 hr modulation of GS1826-24 is confirmed as orbital in origin, this would make it of great interest within the context of transient neutron star LMXB formation and evolution. However, in many respects GS1826-24 does show similarities to other neutron star binaries with comparable periods. The X-ray bursters X1636-536 and X1735-444 exhibit optical modulations of a similar amplitude on their orbital periods (3.80 and 4.65 hr respectively), plus irregular variability on somewhat shorter timescales (\\pcite{whi95} and references therein), although unlike GS1826-24 these systems are not transient on a timescale of decades. The optical emission of LMXBs is dominated by the reprocessed X-rays from the accretion disc and companion. The origin of the underlying sinusoidal modulation is interpreted as the varying contribution from the X-ray heated face of the companion star \\cite{vP95a}. As for the irregular variability, this is probably caused by changes in the spatial distribution of the reprocessing material in the accretion disc and/or fluctuations in the central X-ray luminosity, as suggested in the case of X1636-536 \\cite{vP90b}. Moreover, GS1826-24 has a high $L_X/L_{opt}(\\sim500)$, very similar to that of the compact 41 min binary X1627-673, but lower than the $L_X/L_{opt}\\sim700$ of the 50 min binary X1916-053, which is a higher inclination dipping source (once again these are persistent sources). Since this ratio is related in part to the physical size of the system and the disc reprocessing area available, the similarity supports the hypothesis that GS1826-24 is also a relatively compact system. Furthermore, the non-detection of GS1826-24 prior to 1988 is characteristic of the observed variability of the transient bursting systems X0748-678, X1658-298 and X2129+470 \\cite{whi95}, which have either been detected over many years and then gone into quiescence for a similar period of time or vice-versa. These transients all have known periods in the range 3.82-7.1hrs (see Table \\ref{tab:sptb}). Clearly, further high-speed optical photometric monitoring is required in order to confirm the stability of the 2hr modulation and hence its orbital origin. With only a short {\\it ROSAT} lightcurve published to date \\cite{barr95}, confirmation might be possible from a longer X-ray observation. However, the low-amplitude of the observed modulation ($0.06^m$) implies a low-inclination system ($< 70 \\degree$), and hence X-ray dipping behaviour is unlikely to be seen." }, "9803/astro-ph9803149_arXiv.txt": { "abstract": "The local luminosity function at 25 $\\mu$m provides the basis for interpreting the results of deep mid-infrared surveys planned or in progress with space astrophysics missions including ISO, WIRE and SIRTF. We have selected a sample of 1458 galaxies from the IRAS Faint Source Survey with a flux density limit of 250 mJy at 25 $\\mu$m. The local luminosity function is derived using both parametric and non-parametric maximum-likelihood techniques, and the classical $1/V_{max}$ estimator. Comparison of these results shows that the $1/V_{max}$ estimate of the luminosity function is significantly affected by the Local Supercluster. A maximum-likelihood fit to the radial density shows no systematic increase that would be caused by density evolution of the galaxy population. The density fit is used to correct the $1/V_{max}$ estimate. We also demonstrate the high quality and completeness of our sample by a variety of methods. The luminosity function derived from this sample is compared to previously published estimates, showing the prior estimates to have been strongly affected by the Local Supercluster. Our new luminosity function leads to lower estimates of mid-infrared backgrounds and number counts. ", "introduction": "\\label{sec:intro} Much of the effort to study infrared-luminous galaxies has centered on wavelengths greater than 50 $\\mu$m. Modeling work is focused on the near-IR (e.g. \\markcite{chok94}Chokshi et al. 1994) and far-IR (e.g. \\markcite{hac87}Hacking et al. 1987; \\markcite{rrbinson96}Rowan-Robinson et al. 1996) portions of the galaxian spectrum. However, the mid-infrared is well-suited for studying starburst and ultraluminous galaxies. About 40\\% of the luminosity from starburst galaxies is radiated from 8-40 $\\mu$m (\\markcite{soi87}Soifer et al. 1987). Extinction effects are small, and problems due to infrared cirrus are minimized. Most importantly for space astrophysics, for a fixed telescope aperture, the spatial resolution is higher at shorter wavelengths, and the confusion limit lies at higher redshifts. Recent work using the {\\it Infrared Space Observatory (ISO)} (e.g. \\markcite{knapp96}Knapp et al. 1996; \\markcite{boul96}Boulade et al. 1996; \\markcite{rrobinson96}Rowan-Robinson et al. 1996) shows the relative importance of the 7 $\\mu$m and 15 $\\mu$m bands for galaxy studies. The {\\it Wide-Field Infrared Explorer (WIRE)}, a Small Explorer mission due to launch in late 1998 (\\markcite{hac96}Hacking et al.\\ 1996; \\markcite{schemb96}Schember et al. 1996), will conduct a very deep survey at 24 $\\mu$m to study starburst galaxy evolution. The {\\it Space Infrared Telescope Facility (SIRTF)} is also expected to conduct surveys in mid-infrared bands. To interpret the results of these surveys now in progress or soon to commence, it is necessary to better understand the mid-infrared properties of galaxies in the local Universe. The 25 $\\mu$m luminosity function provides the basis for predicting the faint source counts in the mid-infrared. The empirical model of \\markcite{hac91}Hacking \\& Soifer (1991) uses an analytic fit to the luminosity function derived by Soifer \\& Neugebauer \\markcite{soi91} (1991). This function was estimated from a complete subsample of the Bright Galaxy Sample (\\markcite{soi87}Soifer et al.\\ 1987) containing 135 galaxies to a flux density limit of 1.26 Jy. The availability of many more redshifts of IRAS galaxies (principally from the 1.2 Jy Survey (\\markcite{str90}Strauss et al.\\ 1990; \\markcite{fis95} Fisher et al.\\ 1995)) enables a much larger sample to be studied, reducing uncertainties at high and low luminosities. In this paper we present the selection of a large galaxy sample that is flux-limited at 25 $\\mu$m, and derive the local luminosity function based on this sample. The sample selection is described in the next section. In Section \\ref{sec:lfresults} we describe the $1/V_{max}$ and the maximum-likelihood estimators for deriving the local luminosity function, and present the results. The completeness of the sample is discussed in Section \\ref{sec:compdisc}. In Section \\ref{sec:correct} we calculate the radial density distribution of the sample using a maximum-likelihood method. The radial density fit is used to correct the $1/V_{max}$ estimate of the local luminosity function, as well as the redshift distribution with which the luminosity function can be compared. Section 6 includes discussions of the different luminosity function estimators, a comparison of our newly derived luminosity function with previous estimates, the implications for mid-infrared backgrounds and number counts, and the effects of evolution on the derivation of the luminosity function. The color properties of the sample are treated in another paper (\\markcite{fang98}Fang et al.\\ 1998). ", "conclusions": "\\label{sec:conclusions} The following are the results of this paper: 1. We have selected a sample of 1458 galaxies with redshifts from the IRAS Faint Source Survey with a flux density limit of 250 mJy at 25 $\\mu$m. An additional 17 galaxies do not have redshifts available. 2. The local luminosity function is derived using the $1/V_{max}$ estimator and both parametric and non-parametric maximum likelihood methods. The $1/V_{max}$ estimate is significantly affected by the Local Supercluster. The maximum likelihood methods are independent of density variations, and we consider the parametric fit with parameters in Table \\ref{tab:fitpars} to be the best estimate of the local luminosity function at 25 $\\mu$m. 3. A maximum likelihood fit to the radial density in this sample is used to correct the $1/V_{max}$ estimate. The fit shows no sign of a systematic increase with redshift of the density, as would result from density evolution of the galaxy population. 4. The $1/V_{max}$ luminosity function derived from a smaller sample by Soifer \\& Neugebauer (1991) is significantly contaminated by the Local Supercluster. Predictions of number counts and local luminosity density based on that function are 15-20\\% higher than those indicated by our improved luminosity function. The new function also leads to lower predictions of the mid-infrared background due to galaxies." }, "9803/astro-ph9803180_arXiv.txt": { "abstract": "We report radial velocities for 99 galaxies with projected positions within $30\\arcmin$ of the center of the cluster A3733 obtained with the MEFOS multifiber spectrograph at the 3.6-m ESO telescope. These measurements are combined with 39 redshifts previously published by Stein (1996) to built a collection of 112 galaxy redshifts in the field of A3733, which is used to examine the kinematics and structure of this cluster. We assign cluster membership to 74 galaxies with heliocentric velocities in the interval $10\\,500$--$13\\,000$ \\kms. From this sample of cluster members, we infer a heliocentric systemic velocity for A3733 of $11\\,653^{+74}_{-76}$ \\kms, which implies a mean cosmological redshift of 0.0380, and a velocity dispersion of $614^{+42}_{-30}$ \\kms. The application of statistical substructure tests to a magnitude-limited subset of the latter sample reveals evidence of non-Gaussianity in the distribution of ordered velocities in the form of lighter tails and possible multimodality. Spatial substructure tests do not find, however, any significant clumpiness in the plane of the sky, although the existence of subclustering along the line-of-sight cannot be excluded. ", "introduction": "The rapid development of multifiber spectroscopy in recent years has made possible the simultaneous acquisition of large numbers of galaxy spectra. The obtention of extensive and complete redshift data bases for clusters of galaxies has hastened the investigation of the physical properties of their visual component which, in turn, is allowing for a better understanding of the characteristics of the dark matter distribution on Mpc scales. Here, we report a total of 104 redshift measurements for 99 galaxies in the field of A3733 and use these data, in combination with a previously published sample of 39 redshifts, to perform a kinematic and spatial analysis of the central regions of this cluster. A3733 is a southern galaxy cluster listed in the ACO catalog (Abell, Corwin, \\& Olowin 1989)\\cite{ACO89} as of intermediate Abell's morphological type and richness class $R=1$. This cluster hosts a central cD galaxy, included in the Wall \\& Peacock (1985)\\cite{WP85} all-sky catalog of brightest extragalactic radio sources at 2.7 GHz, which has led to its classification as of Bautz-Morgan type I--II (Bautz \\& Morgan 1970)\\cite{BM70}. A3733 is also one of the 107 nearby rich ACO clusters ($R\\ge 1$, $z\\le 0.1$) included by Katgert et al. (1996)\\cite{Ka96} in the ESO Nearby Cluster Survey (ENACS), as well as a one of the X-ray-brightest Abell clusters detected in the ROSAT All-Sky Survey by Ebeling et al. (1996)\\cite{Eb96}. The only major kinematical study of A3733 done so far is that of Stein (1997)\\cite{St97}. From a sample of 27 cluster members located within $r\\la 16\\arcmin$ from the cluster center, this author has found no evidence of significant substructure in the cluster core. This study of A3733, which is part of a more general investigation of the frequency of substructure in the cluster cores from an optical spectroscopic survey conducted on a sample of 15 nearby ($0.01\\la z\\la 0.05$) galaxy clusters (Stein 1996)\\cite{St96}, is based on a dataset that has many characteristics in common with the ENACS data gathered for the same field. Indeed, the two datasets have been obtained with the OPTOPUS multifiber spectrograph at the ESO 3.6-m telescope and cover essentially the same area on the sky. Besides, they have also a very similar number of galaxies: 39 and 44, respectively (28 of which are shared). The MEFOS redshift dataset for A3733 reported in this paper contains two and a half times the number of galaxy radial velocities reported by Stein (1996)\\cite{St96}, including 26 reobservations, while it covers a circular region around the center of A3733 four times larger. Furthermore, its high degree of completeness offers the possibility of extracting a complete magnitude-limited subset with a number of galaxies large enough for its use on statistical analysis. The plan of the paper is as follows. In Sect.~2 we discuss the MEFOS spectroscopic observations and data reduction, and present a final sample with 112 entries built by the combination of the MEFOS and Stein's (1996)\\cite{St96} data. Section~3 begins with a brief description of the tools which will be used for the analysis of the data. Next, we identify the galaxies in our sample that belong to A3733, and use this dataset and a nearly complete magnitude-limited subset of it to examine the kinematical properties and structure of the central regions of the cluster. Section~4 summarizes the results of our study. ", "conclusions": "We have reported 104 radial velocity measurements performed with the MEFOS multifiber spectrograph at the 3.6-m ESO telescope for 99 galaxies in a region of $30\\arcmin$ around the center of the cluster A3733. To augment this data, we have combined the MEFOS measurements with 39 redshifts measured by Stein (1996)\\cite{St96} with the OPTOPUS instrument at the same telescope. This has given a final dataset with a total of 112 entries in the field of A3733. Radial velocities have been then supplemented by COSMOS \\bj\\ magnitudes and accurate sky positions in order to investigate the kinematics and structure of the central regions of the cluster. From a sample containing 74 strict cluster members, we have derived a heliocentric systemic velocity for A3733 of $11\\,653^{+74}_{-76}$ \\kms, resulting in a $\\overline{z}_{\\mathrm CMB}$ of 0.0380, and a velocity dispersion of $614^{+42}_{-30}$ \\kms, in good agreement with the estimates by Stein (1997)\\cite{St97} from the OPTOPUS data alone. Statistical tests relying exclusively on the distribution of observed velocities have yield suggestive indication of the possible kinematical complexity of A3733, especially when applied to a nearly complete magnitude-limited (\\bj\\ $\\leq 18$) sample of cluster members. Despite this result, two powerful substructure tests that incorporate spatial information have failed to detect in this latter sample any statistically significant evidence of clumpiness in the galaxy component, in agreement with the findings of a previous study based on a spatially less extended and less complete dataset. Given that the sensitivity of the spatial substructure tests we have used is reduced when the subunits are seen with small projected separations, the results of the present study cannot exclude, however, the possibility that the signs of kinematical complexity detected in the velocity histogram of A3733 might be due to the existence of galaxy subcondensations superposed along the line-of-sight." }, "9803/astro-ph9803194_arXiv.txt": { "abstract": "An analysis of the observed characteristics of the Galactic Cepheid variables is carried out in the framework of their \\plr\\ being used as a standard candle for the distance measurement. The variation of the observed number density of Galactic Cepheids as function of their period and amplitude along with stellar pulsation characteristics is used to divide the population into two groups: one with low periods, probably multi-mode or higher mode oscillators, and another of high period variables which should be dominantly fundamental mode radial pulsators. Methods to obtain extinction-corrected colors from multi-wavelength observations of the second group of variables are described and templates of the $\\vi $ \\lig s are obtained from the $V$ \\lig s. Colors computed from the model atmospheres are compared with the extinction-corrected colors to determine the Cepheid instability strip in the {\\em mean surface gravity--effective temperature diagram}, and relations are derived between mean colors {\\em $\\BV$ vs period of pulsation}, {\\em $\\vi$ vs period}, and {\\em $\\vi$ at the brightest phase vs amplitude of pulsation}. The strength of the $\\kappa$-mechanism in the envelope models is used to estimate the metal dependency of the instability strip from which an idea of the sensitivity of the \\plr\\ to the helium and metal abundance is given. Some estimate of the mass of Cepheids along the instability strip is provided. ", "introduction": "\\label{sec:intro} The classical Cepheid variables provide an important standard candle to measure distances to galaxies up to $\\sim 30$ Mpc. The Cepheid Distance Scale is considered to be among the most reliable methods because the physics of Cepheid pulsation is well-understood and the relation between the pulsation period and luminosity of the star is observationally well-established. The Cepheids are luminous, have a narrow range of surface temperatures; their pulsation is very stable and exhibit large amplitude. The intrinsic scatter in their \\plr\\ is believed to be only around $0.3$ mag. However, the Cepheid distance scale cannot be directly calibrated from the observation of nearby stars and consequently, several systematic effects still undermine its effectiveness as a standard primary candle to determine extragalactic distances beyond a few Mpc. Some of the questions which have direct bearing on the problem of distance calibration, but whose answers remain inconclusive in spite of extensive research, are listed below. \\bei \\item Are the preferential pulsation modes of Cepheids period-dependent? \\item Is a single \\plr\\ applicable to the entire instability strip? \\item Is the \\plr\\ modified appreciably due to metallicity dependency of the stellar structure? \\item Is the \\plc\\ relation a better indicator of distance than the \\plr\\ only? \\eni \\noindent Iben and Tuggle (1975) numerically computed the period and luminosity of Cepheids for a range of masses and obtained a relation between metallicity, surface temperature, period and luminosity. The \\plc\\ relation is found to be dependent on metallicity due to extreme sensitivity of the color--temperature relation on chemical composition. However, according to Becker, Iben and Tuggle (1977), within the uncertainties, the relation between period and luminosity for the first and second crossings of the Cepheid instability strip by a particular star does not crucially depend on the chemical composition. Since the time spent in traversing the strip is largest for the second crossing, most of the observed Cepheids are in this stage of evolution. So, although conversion of the period into a $V$-magnitude will introduce small effects due to surface temperature and metallicity, a \\plr\\ derived from observations should not be affected by small variations in the chemical composition. Indeed, the robustness of the \\plr\\ against changes in the chemical composition is borne out by theoretical as well as observational studies. The theoretical models of Bressan \\ea (1993) produce the same period (within an error of $2 \\% $) for a given luminosity, irrespective of the chemical composition. From observations of Cepheids in M31, it appears that there is no significant dependence of the \\pl\\ zero point on metallicity gradients (\\cite{fm:90}). The recent review on the metallicity dependence of the Cepheid Distance Scale in the context of the HST Key Project on Extragalactic Distance Scale (\\cite{kenn:98}) also leads to the same conclusion. Similarly, even though color-color diagram of Cepheids can be used to determine their metallicity, it is not an improvement in terms of its application to the estimation of distances, given that after extinction correction the color has larger error than the $V$-magnitude (e.g., \\cite{fg:93}). Clearly, in order to address the above question, systematic work is required on the pulsation properties, evolution as well as stellar atmospheric structure, taking into account the onset of convection in the atmosphere during the pulsation cycle. In the present study we shall adopt the working hypothesis that the luminosity of the star {\\it as a function of period} is not directly altered by metallicity (subject to the star being a classical Cepheid), and that the color of the star provides a better diagnostic for the estimation of extinction than for the determination of the distance. We devise methods to determine the extinction by the interstellar medium, and particularly emphasize the importance of observations in multi-wavelength bands, and also address the question of pulsation modes of Cepheids. The Cepheids in our own Milky Way galaxy have been observed by several astronomers over the years, and it is possible to obtain multi-wavelength data as well as accurate periods for a large number of them. A careful analysis of these Galactic Cepheids will naturally provide a useful template for identifying and estimating the various errors in the calibration of the Cepheid Distance Scale. A robust calibration of this distance scale is particularly important for extending it to extragalactic domains, as Cepheids are being observed in several far away galaxies, including those in the Virgo Cluster, by the Hubble Space Telescope. The hope is that distances based on these observations will ultimately lead to an accurate determination of the Hubble Constant. In this context, we attempt to provide a new calibration of the Cepheid Distance Scale, which is free from many of the systematic errors. In an accompanying communication (which we will refer to as Paper II), we apply these results for estimating the distance to the Virgo Cluster, based on the HST data for the Cepheids in the spiral M100. This paper is organized as follows. In Section~\\ref{sec:number} we discuss the number distribution of Cepheids against their periods. In Section~\\ref{sec:ligcur}, we demonstrate the feasibility of obtaining accurate $\\vi $ \\lig s from the $V$ \\lig\\ of a Cepheid and limited number of observations in the I band. This method is particularly useful for analyzing Cepheid data with few observations in one band (as in HST observations). In Section~\\ref{sec:extcor}, we devise a formalism for extinction correction for each individual Cepheid, based on model atmospheres and an $\\rv$-dependent extinction law. Several useful period--color and amplitude--color relationships are also derived. Section~\\ref{sec:modes} is concerned about the different modes of pulsation of Cepheid variables and their manifestations in the observed properties like period and amplitude of pulsation. We argue that it is necessary to choose the correct lower cutoff period in the \\plr\\ in order to prevent contamination from multi-mode pulsators. In Section~\\ref{sec:mass}, we give an estimate for Cepheid masses at different periods, based on our results about the instability strip in the surface gravity versus effective temperature plane. Some discussion on the metallicity effects are presented in Section~\\ref{sec:metal} and the major limitations of the present work are listed in Section~\\ref{sec:limit}. The main conclusions from this work are summarized in Section~\\ref{sec:concl}. ", "conclusions": "" }, "9803/hep-ph9803412_arXiv.txt": { "abstract": "We show that the decaying magnetohydrodynamic turbulence leads to a more rapid growth of the correlation length of a primordial magnetic field than that caused by the expansion of the Universe. As an example, we consider the magnetic fields created during the electroweak phase transition. The expansion of the universe alone would yield a correlation length at the present epoch of 1 AU, whereas we find that the correlation length is likely of order 100 AU, and cannot possibly be longer than $10^4$ AU for non-helical fields. If the primordial field is strongly helical, the correlation length can be much larger, but we show that even in this case it cannot exceed 100 pc. All these estimates make it hard to believe that the observed galactic magnetic fields can result from the amplification of seed fields generated at the electroweak phase transition by the standard galactic dynamo. ", "introduction": "Recently, considerable interest has been focused on the possibility that a primordial magnetic field may have been created at some early stage of the evolution of the Universe. While the existence of a weak widespread extragalactic chaotic magnetic field cannot be ruled out, almost the only place where such a primordial field may have left an observable imprint is in galaxies, many of which possess microgauss, kiloparsec scale fields that are thought to be the result of dynamo amplification of a weak seed field \\cite{Kronberg}. The idea that the seed for the galactic dynamo may be the field created in the very early Universe has inspired a number of works. For example, such fields may have appeared at the electroweak phase transition \\cite{BaymMcLerran}. There are numerous other proposals involving physics at various scales (for a brief overview and further references, see \\cite{Olesen-rev}). One feature shared by most particle-physics scenarios is the smallness of the correlation length of the magnetic field which results. Indeed, at the moment of creation, the correlation length is limited by the horizon radius. By pushing this moment to a very early stage in the history of the Universe, one makes the correlation length much smaller than it would be if the magnetic fields were created more recently, say, during proto-galactic contraction \\cite{Ostriker}. For example, the fields generated at the electroweak phase transition, when the horizon radius is about 1 cm, would have a correlation length of at most about $10^{15}$ cm in the present Universe. Contraction of proto-galaxies is likely to reduce this length scale by 2 orders of magnitude, which gives $10^{13}$ cm, or 1 AU, as the characteristic length scale of the seed field which the galactic dynamo is supposed to amplify. Realistically, this length scale is likely much smaller, since the correlation length at creation is typically much less than the horizon size\\footnote{A magnetic field created during or before inflation \\cite{TurnerWidrow} may have a large correlation length, however the proposed inflationary models that could possibly generate a large-scale magnetic field may seem rather contrived \\cite{Ratra}. We will not consider inflation-induced magnetic fields in this paper, but one should keep in mind this alternative.}. Because any primordial magnetic fields generated at early times (via particle physics occurring at, say, the electroweak scale) have such small correlation lengths, these fields are not attractive candidates to play the role of seed fields for the galactic dynamo, unless the correlation length is somehow increased. The best developed theory of magnetic amplification in galaxies, the mean-field dynamo \\cite{Ruzmaikin}, operates under the assumption that the magnetic field is smooth on the scale of turbulent motion of the interstellar gas, which is of order 100 pc. One may attempt to apply the mean-field dynamo theory for large-scale Fourier components of the chaotic magnetic field, neglecting the small-scale ones, however it is not obviously the correct procedure. That the magnetic field at very small scales exponentiates rapidly is a well known fact that poses a serious problem for the mean-field dynamo theory \\cite{KurlsrudAnderson}. In the situation when the seed field itself resides at small scales, this problem is likely to become more severe. In this paper, we will not address the question of how a seed field is created, whether at the electroweak phase transition or during some other early epoch. Our goal is to investigate the possibility that magnetohydrodynamic (MHD) effects can lead to a substantial increase in the length scale of the magnetic field at the present epoch. We show that decaying MHD turbulence typically leads to a faster growth of the magnetic correlation length than one would expect from the expansion of the Universe alone. We estimate that in the case of magnetic fields generated at the electroweak phase transition, the enhancement factor is $10^2$, and the correlation length may reach 100 AU. If the primordial field has a large Chern-Simons number, the enhancement factor may be much larger, but the correlation length cannot possibly exceed 100 pc today if the magnetic field comes from the electroweak epoch. Thus, although the magnetohydrodynamic effects we consider are of some help, they cannot increase the correlation length enough to make the electroweak generation of a primordial seed field a viable option. We also consider generating the seed field at the QCD phase transition. This may be a possibility, but is only viable if the bubble separation at the phase transition is very large. This paper is organized as follows. In Sec.\\ \\ref{sec:MHDeq} we write the basic equation governing the MHD of the Universe. Sec.\\ \\ref{sec:non-hel} is devoted to the decay of non-helical MHD turbulence, whereas Sec.\\ \\ref{sec:hel} describes the scaling laws of the decay of helical turbulence. Sec.\\ \\ref{sec:concl} contains concluding remarks. The Appendix contains details about the EDQNM approximation used in our numerical simulation. ", "conclusions": "\\label{sec:concl} In this paper we have considered the behavior of the correlation length of the primordial magnetic field, if such a field is generated in the early Universe. We found the decay law of the MHD turbulence, which is responsible for a substantial increase of the correlation length. We also observed a qualitative difference between the cases of non-helical and helical initial magnetic field. Taking the case in which the magnetic field is generated during the electroweak phase transition as an example, we find that if the magnetic field is not helical the factor gained from MHD is about $10^2$ and today the field may be correlated at scales as large as 100 AU. If by some chance the field is helical, the enhancement factor is much larger, but the final correlation length cannot exceed 100 pc, corresponding to 1 pc after proto-galactic collapse. Let us note some uncertainties remaining in our estimation. (In general, addressing these issues would lead to less optimistic values of the enhancement factor.) First, it is not at all clear whether in MHD turbulence the magnetic energy reaches equipartition with the kinetic energy of bulk fluid motion. While semi-analytical calculations (including our simulation) favor equipartition \\cite{Pouquet}, some numerical results indicate that the mean magnetic energy density remains small and concentrated at scales shorter than the largest turbulence scale \\cite{dns}. If the latter remains true at very high Reynolds number, the magnetic field will be smaller after the electroweak phase transition and the correlation length at present will be smaller than we have estimated. The second factor is neutrino diffusion. After the electroweak phase transition, neutrinos are the particles with the longest mean free path. There is a certain time interval when neutrinos are still not decoupled from the physics at the turbulence scale, but their diffusion length is so large that the neutrino contribution to viscosity makes the Reynolds number to drop below 1. During this time, there is no turbulence and the magnetic field is frozen. After the neutrinos decouple from the fluid motion at the turbulence scale, magnetic stress leads to restoration of turbulence, and the magnetic length may continue to grow. The overall effect is some reduction in the estimate of the final magnetic correlation length. It would be nice to end this paper on a more positive note, and to this end we turn now to investigating the QCD phase transition, which occurs later than the electroweak transition and so may yield a longer correlation length. If the QCD phase transition is first order, it will introduce fresh turbulence stronger than the decaying electroweak turbulence. The amplitude and correlation length of the magnetic field will be determined by the QCD turbulence. If fields were regenerated at the QCD transition on length scales of order of the horizon, the length scale it would have today is very large, since the horizon size at the QCD phase transition is much larger than at the electroweak epoch. Even with no enhancement from MHD effects, the expansion of the universe yields a magnetic correlation length of order 1 pc. It can be estimated that the turbulence will survive till at least matter-radiation decoupling, at which the correlation length is is of order 1 kpc in the non-helical case and 100 kpc if the field is helical. However, a magnetic field correlated on the horizon size is not expected to be produced at the QCD phase transition, and these estimates are too optimistic. Instead, the initial length scale must be of the order of the bubble spacing at the end of the phase transition (which is the natural scale of turbulence). Bubble spacings larger than 100 cm are unlikely \\cite{Ignatius} and may have problem with standard big bang nucleosynthesis \\cite{bbn}. A bubble spacing smaller than 100 cm again implies a very small correlation length at the present epoch. If this constraint is respected, our analysis of the QCD case must end pessimistically, as in the electroweak case. However, one cannot completely rule out the possibility of a very large bubble spacing, which leads to a magnetic field correlated on a long enough length scale to be of interest. A nonstandard evolution in which mixing after the phase transition occurs rapidly by hydrodynamic flows instead of slowly by diffusion or nonstandard nucleosynthesis may be required in this case. Finally, although in this paper we presented the smallness of the magnetic correlation length as undesirable and tried to overcome it by invoking MHD turbulence, there may exist a non-standard dynamo mechanism where a small-scale seed gives rise to a large-scale magnetic field (see, e.g., \\cite{Chandran}). This possibility is, however, outside the scope of the present paper." }, "9803/astro-ph9803257_arXiv.txt": { "abstract": "We present a systematic search for OVI(1032\\AA,1037\\AA) absorption in a Keck HIRES spectrum of the $z=3.62$ quasar Q1422+231, with the goal of constraining the metallicity and ionization state of the low density intergalactic medium (IGM). Comparison of CIV absorption measurements to models of the \\lya forest based on cosmological simulations shows that absorbers with $\\nh \\ga 10^{14.5}\\cdunits$ have a mean carbon abundance [C/H]~$\\approx -2.5$, assuming a metagalactic photoionizing background with the spectral shape predicted by Haardt \\& Madau (1996, HM). In these models, lower column density absorption arises in lower density gas where most CIV is photoionized to CV. Therefore, OVI should be the most sensitive tracer of metallicity in \\lya absorbers with $\\nh \\la 10^{14.5}\\cdunits$. OVI lines lie at wavelengths heavily contaminated by Lyman series absorption, so we interpret the search results by comparing to carefully constructed, mock Q1422 spectra drawn from a hydrodynamic simulation of a $\\Lambda$-dominated cold dark matter model. A search for deep, narrow absorption features yields only a few candidate OVI lines in the spectrum of Q1422. HI absorption blankets the position of the doublet companion line in each case, and the total number of narrow lines is statistically consistent with that in zero-metallicity artificial spectra. Artificial spectra generated with the HM background and [O/H]~$\\ga -2.5$ predict too many narrow lines and are statistically inconsistent with the data. We also search for OVI associated with CIV systems, using the optical depth ratio technique of Songaila (1998). With this method we {\\it do} find significant OVI absorption; matching the data requires [O/C]~$\\approx +0.5$ and corresponding [O/H]~$\\approx -2.0$. Taken together, the narrow line and optical depth ratio results imply that (a) the metallicity in the low density regions of the IGM is at least a factor of three below that in the overdense regions where CIV absorption is detectable, and (b) oxygen is overabundant in the CIV regions, consistent with the predictions of Type II supernova enrichment models and the observed abundance pattern in old halo stars. The photoionizing background spectrum would be truncated above 4~Ry in regions that have not undergone helium reionization (HeII$\\longrightarrow$HeIII), and in this case matching the Q1422 data requires lower [C/H] but higher [O/H]. Taking [O/C]$\\approx +1$ as the maximum plausible overabundance of oxygen, we conclude that helium must have been reionized through at least 50\\% of the volume from $z \\sim 3 - 3.6$. ", "introduction": "The Lyman alpha (Ly$\\alpha$) ``forest\" (\\cite{lyn71}; \\cite{sar80}) of spectral features caused by HI absorption along the line of sight to a quasar probes the state of the intergalactic medium over a wide range of physical conditions. During the past few years, high-precision observations made using the HIRES spectrograph (\\cite{vog94}) on the 10m Keck telescope have quantified the statistics of these low column density absorbers to unprecedented accuracy (e.g., \\cite{hu95}; \\cite{lu96b}; \\cite{kim97}; \\cite{kir97}). During the same time span, cosmological simulations that incorporate gas dynamics, radiative cooling, and photoionization have been able to reproduce many of the observed properties of quasar absorption spectra (\\cite{cen94}; \\cite{zha95}; \\cite{her96}; \\cite{mir96}; \\cite{dav97a}). The rapid progress on theoretical and observational fronts has led to the emergence of a new paradigm for the origin of the high-redshift ($z\\ga 2$) \\lya forest, in which most \\lya forest lines are produced by regions of low to moderate overdensity in hierarchically collapsing structures that are not in dynamical or thermal equilibrium. \\lya lines of lower column density generally arise in gas of lower physical density, which has a lower neutral hydrogen fraction because of the reduced recombination rate. In this paper, we use a matched comparison between a cosmological hydrodynamic simulation and a Keck HIRES spectrum of the quasar Q1422+231 to constrain the metal abundance of this low density gas. The recent detection of metal lines associated with \\lya forest absorbers having column densities $\\nh \\la 10^{15} \\cdunits$ (\\cite{cow95}; \\cite{tyt95}; Songaila \\& Cowie 1996, hereafter \\cite{son96}) has provided a new avenue for investigating the ionization state and enrichment history of the high-redshift intergalactic medium (IGM). \\cite{son96} showed that 75\\% of \\lya absorbers with $\\nh > 10^{14.5} \\cdunits$ have associated CIV absorption. Using simple photoionization models they estimate that the mean metallicity of these absorbers is between\\footnote{We use the standard notation of brackets to denote the relative abundance, in logarithm, versus solar.} [C/H]$\\sim -2$ and $-3$. Studies that use cosmological simulations to model the density and temperature of the absorbing gas obtain a better match to the data for a mean metallicity of [C/H]$\\sim -2.5$ with around one dex of scatter, assuming either a power-law ionizing background (\\cite{hae96}) or a reprocessed quasar ionizing background (Hellsten \\etal 1997, hereafter \\cite{hel97}). The question of how metals came to reside in these intermediate-density \\lya absorbers remains unanswered. Since \\lya absorbers with $\\nh \\la 10^{15}\\cdunits$ are optically thin and associated with density peaks of overdensity $\\la 20$, it is unlikely that they contain star-forming regions that produce {\\it in situ} enrichment. Thus some transport mechanism must be invoked to explain the presence of metals in these regions. One possibility is that the metals are ejected from nearby, forming galaxies, by some combination of supernova blowout (\\cite{mir97}) and tidal stripping (\\cite{gne97}; \\cite{gne98}). Since the enriching proto-galaxies are likely to form more efficiently in high density environments, these scenarios predict a significant correlation between metallicity and density (see, e.g., figure~6 of Gnedin \\& Ostriker 1997). Alternatively, the IGM may have been enriched at a very early epoch by more ubiquitous Population III objects (e.g., \\cite{hai97}), in which case all \\lya absorbers might be expected to have roughly the same metallicity. Because CIV lines probe only a small range of HI column densities (and hence physical densities) with adequate statistics, it is difficult to distinguish between these enrichment models using only CIV data. Metal abundances can be studied at higher densities in Lyman limit systems ($\\nh \\ga 10^{17}\\cdunits$) and damped \\lya systems ($\\nh \\ga 10^{20}\\cdunits$), but these are objects where {\\it in situ} enrichment is likely and where (especially for Lyman limit systems) uncertain radiative transfer effects complicate the inference of metal abundances from line strengths. In this paper, we attempt to extend metallicity constraints to the low column density ($\\nh \\la 10^{14.5}\\cdunits$), and hence low physical density, \\lya forest. Carbon is difficult to detect in this regime because the low density reduces $N_{C}$ and ionizes more CIV to CV. However, \\cite{lu98} have used composite spectra to attempt to detect CIV in such systems, and we briefly discuss their results in \\S\\ref{sec: disc}. In this paper, we focus instead on OVI, which is expected to be the one detectable metal absorption feature tracing \\lya absorbers with $\\nh \\la 10^{14.5}\\cdunits$ because of its high ionization state and large oscillator strength (Hellsten \\etal 1998, hereafter \\cite{hel98}). The difficulty with this approach, and the reason that it has not been previously attempted, is that the OVI absorption features lie embedded within the \\lya forest, which is quite crowded at these redshifts. We overcome this problem by using a line identification scheme specifically designed to select candidate OVI features and by using artificial spectra extracted from a realistic cosmological simulation to calibrate the efficiency of OVI detection and the contamination from narrow \\lya lines. With these procedures, we can test whether the low density regions of the \\lya forest are consistent with a uniform metallicity extrapolated from the CIV data at higher densities. The low density IGM is highly photoionized by the metagalactic ultraviolet (UV) background. For our standard spectral shape, we assume that the UV background is produced by quasar emission reprocessed by \\lya forest absorption (Haardt \\& Madau 1996, hereafter \\cite{haa96}). We find that if the IGM metallicity is uniform at [C/H]$\\sim -2.5$, the UV background has the spectral shape given by \\cite{haa96}, and oxygen has the factor of three overabundance (relative to solar) predicted by Type II supernova enrichment models, then OVI should be readily detectable in the spectrum of Q1422+231. However, our detection algorithm finds very few candidate OVI lines in the spectrum. The absence of detectable OVI features has several possible interpretations: (1) oxygen is not overabundant relative to carbon in the low density IGM, (2) the metallicity of low density regions as traced by $\\nh \\la 10^{14.5}\\cdunits$ \\lya absorption systems is lower than the metallicity of intermediate-density regions traced by higher column density absorption, or (3) there are many fewer high-energy photons capable of photoionizing OV to OVI than are predicted by the \\cite{haa96} ionizing background. To distinguish between these interpretations, we apply a second algorithm, the optical depth ratio technique of Songaila (1998, hereafter \\cite{son98}) designed to detect OVI in regions where significant CIV absorption is found. Since we have independently determined [C/H] in these regions, we can use this technique to discriminate between different ionization conditions and abundance patterns. By applying this technique to the spectrum of Q1422+231 and calibrating the results using artificial spectra, we find that: (1) a significant metallicity gradient (declining metallicity with declining density) must exist regardless of whether helium has mostly reionized or not, (2) if helium has reionized by $z\\sim 3.6$, our results are consistent with [O/C]~$\\approx +0.5$, and (3) if the epoch of helium reionization does not begin until $z\\sim 3$, a highly implausible oxygen overabundance of [O/C]~$\\ga +2$ is required. Combining the results from these two analysis techniques, our favored scenario for the low density IGM at $z\\ga 3$ is one in which more than half of the volume of the universe has helium predominantly reionized by $z\\sim 3.6$, oxygen is overabundant relative to carbon by a factor $\\ga 3$, and spatial regions with overdensities $\\sim 10$ have an average metallicity {\\it at least} a factor of 3 higher than regions near the mean baryonic density. Section~\\ref{sec: modeling} describes our cosmological simulation, the Q1422+231 data, and our procedure for constructing artificial absorption spectra. Section~\\ref{sec: OVIsearch} discusses previous OVI searches and reviews \\cite{hel98}'s argument that OVI should be the most effective tracer of metallicity in \\lya absorbers with $\\nh \\la 10^{14}\\cdunits$. Section~\\ref{sec: civ} examines CIV absorption at intermediate column densities, repeating the general arguments of \\cite{hel97} and Rauch et al.\\ (1997a) but with a much closer match between the theoretical and observational analysis procedures. Section~\\ref{sec: search} is the heart of the paper. It describes our algorithm for identifying candidate OVI lines, presents the results of the OVI searches in the real and artificial spectra, and discusses the properties of candidate OVI absorbers in Q1422+231 and the simulations. Figures~\\ref{fig: auto5}--\\ref{fig: OVIsel} demonstrate the paper's central results. Section~\\ref{sec: assumptions} discusses the impact of varying the assumptions of our standard theoretical model. Section~\\ref{sec: pixmet} describes the optical depth ratio technique for quantifying OVI absorption, shows the results from this technique applied to Q1422+231 and artificial spectra, and discusses these results in conjunction with the results from Section~\\ref{sec: search}. Section~\\ref{sec: disc} summarizes our conclusions and discusses them in light of other recent observational and theoretical developments. ", "conclusions": "\\label{sec: disc} We present a systematic search for OVI absorption in the spectrum of Q1422+231 ($z=3.62$), using a narrow line detection algorithm proven effective at identifying OVI absorption in artificial spectra, and an optical depth ratio technique introduced by \\cite{son98}. The first technique traces OVI predominantly in systems with $10^{13.5}\\la \\nh\\la 10^{15}\\cdunits$, whereas the second technique traces only OVI associated with CIV absorption, \\ie in $10^{14.5}\\la \\nh\\la 10^{16}\\cdunits$ systems. By comparing Q1422 and artificial spectra having varying metallicities, we determine that \\begin{enumerate} \\item{[O/H] must be lower in lower density regions, for either an \\cite{haa96} ionizing background or an ionizing background significantly truncated above 4~Ry. If [O/C] is constant in systems up to $\\nh\\sim 10^{16}\\cdunits$, our results imply that regions traced by $\\nh\\la 10^{14}\\cdunits$ systems (corresponding to gas at roughly the mean baryonic density) have a mean metallicity lower by at least a factor of 3 compared to regions traced by $\\nh\\sim 10^{15}\\cdunits$ systems (corresponding to a baryonic overdensity of $\\sim 10$).} \\item{More than half the universe must have helium reionized by $z\\sim 3$. If helium has completely reionized by $z\\sim 3.6$ (the highest redshift probed by the Q1422 data), then our analysis implies [O/C]~$\\approx +0.5$, in good agreement with overabundance measurements of Type II supernovae enriched systems. If a significant portion of the universe has not reionized helium by $z\\sim 3$ and therefore has a softer ionizing background spectrum, then the required oxygen overabundance is higher. For example, if half of the volume has not reionized helium, then [O/C]~$\\approx +1.2$, already greater than the observed overabundance of any class of Type II supernovae enriched systems. If the spectrum were soft throughout the universe at $z\\ga 3$ then an implausibly high overabundance, [O/C]~$\\approx +2.3$, would be required. } \\end{enumerate} These conclusions are in good agreement with the recent study of \\cite{lu98}, who used composite spectra to investigate CIV absorption in systems with $10^{13.5}< \\nh < 10^{14} \\cdunits$ and found that the metallicity of these absorbers must be [C/H]~$\\la -3.5$. Cosmological simulations show that $\\nh \\sim 10^{14} \\cdunits$ roughly corresponds to the dividing line between overdense and underdense regions of the universe (though the value of $\\nh$ that marks this division depends on redshift and, to a lesser extent, on cosmological parameters). Our results therefore imply that mildly overdense regions such as filaments and sheets have been enriched, while underdense regions are virtually chemically pristine. Simulations that self-consistently enrich the IGM by tracking metal production and transport find that a strong metallicity gradient is predicted between the mildly overdense and underdense regions (see figure~3 in \\cite{gne98}); this predicted gradient is in good agreement with the \\cite{lu98} data and with the scenario we present above. Recent measurements of a jump in the SiIV/CIV ratio around $z\\sim 3$ (\\cite{son98}) may be difficult to reconcile with conclusion~(2) above. While we have yet to conduct a systematic comparison of SiIV in observed and artificial spectra, primarily because of the small numbers of SiIV systems detectable in our one available quasar spectrum, we expect our results will be in agreement with \\cite{son98}, who argues that such a jump requires a much softer ionizing background at $z\\ga 3$. One way to reconcile these results may be to invoke patchy helium reionization at that epoch, as suggested by \\cite{rei97}; such a model may be tested in greater detail by searching for an anti-correlation between OVI and SiIV detections. If helium reionization occurs around $z\\sim 3$, our narrow line algorithm should yield many more OVI detections at redshifts $z\\la 3$. Such searches are difficult because of the poor blue sensitivity of the HIRES spectrograph (and complete loss of sensitivity at $\\lambda\\la 3800$\\AA), but quasar spectra do exist that could provide constraints down to $z\\sim 2.7$. The presence of a substantial number of OVI lines in this regime would strongly favor the late helium reionization scenario; there is already some weak evidence that OVI is more abundant at $z\\la 3$ (\\cite{son98}). If OVI features continue to be virtually undetectable down to $z\\sim 2.7$, this would be compelling evidence against the late helium reionization scenario, since the HeII absorption measurements of Davidsen, Kriss, \\& Zheng (1996) imply that helium has been reionized by this redshift. We hope to work with observers to attempt this search in the near future. In a broader context, our work illustrates the power of combining cosmological hydrodynamic simulations of structure formation with high-quality quasar spectra to infer the ionization state and the enrichment history of the high-redshift IGM. Future observations and simulations promise a wealth of information, which, when combined, will help us to better understand the evolution of the IGM and its connection to early star formation and the epoch of primeval galaxies." }, "9803/gr-qc9803026_arXiv.txt": { "abstract": "Clues as to the geometry of the universe are encoded in the cosmic background radiation. Hot and cold spots in the primordial radiation may be randomly distributed in an infinite universe while in a universe with compact topology distinctive patterns can be generated. With improved vision, we could actually see if the universe is wrapped into a hexagonal prism or a hyperbolic horn. We discuss the search for such geometric patterns in predictive maps of the microwave sky. ", "introduction": "\\label{compf} We want to predict a map of the temperature fluctuations. In a homogeneous and isotropic space, an angular average over the fluctuations contains all of the essential information. Of the six compact, orientable flat spaces, all destroy global isotropy and all except for the hypertorus destroy global homogeneity. As a result, there is more information in a map of temperature fluctuations than just the angular power spectrum. Although we argue that the angular average overlooks conspicuous features in general, for the equal sided flat cases angular spectra do provide a reasonable bound. Four of the six orientable, compact topologies of $\\e$ are constructed from a parallelepiped as the fundamental domain. The other two are built from a hexagonal prism. The hypertorus is the simplest and has been studied by many authors \\cite{{flat},{moreflat},{sss},{add}}. Stevens, Scott, and Silk \\cite{sss} pointed out that in a flat $3$-torus, the spectrum of temperature fluctuations was truncated at long wavelengths in order to fit within the finite box. Contrary to standard lore, we find all of the equal sided compact flat manifolds show a truncation in the power of fluctuations on wavelengths comparable to the size of the fundamental domain \\cite{{lss},{tarun}}. The longest wavelength fluctuation observed, namely the quadrupole, is in fact low. Some might even take this as evidence for topology \\cite{workshop}. Cosmic variance is also large on large scales. Consequently, a fundamental domain the size of the observable universe is actually consistent with the COBE data \\cite{lss}. A very small universe however is incompatible with the data. The cutoff in long wavelength perturbations is accompanied by gaps in power at wavelengths that do not correspond to integer windings through the fundamental domain. {\\it All} compact spaces show discrete harmonics and as such the sharp harmonics may be a more generic sign of compact topology. The jaggy spectra of such small compact flat spaces are tens of times less likely than the smooth spectrum of infinite $\\e$. We conclude, quite conservatively, that the universe, if finite and flat and equal-sided, must be at least $80$\\% the radius of the surface of last scatter and so $40$\\% of the diameter of the observable universe. There could still be as many as eight copies of our universe within the observable horizon. \\begin{figure} \\centerline{{\\psfig{file=hexglue.eps,width=3in}}} \\centerline{{\\psfig{file=hex0_1_.1.ps,width=3in}}} \\caption{A hexagonal prism with a $2\\pi/3$ twist. The observer is at the center of the universe in the map of $\\delta T(\\hat n)/T$. The fundamental domain is half the diameter of the observable universe in two directions and one-tenth that in the twisted direction. \\label{fighex}} \\end{figure} If instead of an equal-sided space we consider a fundamental domain with disparate length scales, the angular power spectrum is in general a poor discriminant. The averaging over the sky fails to recognize the strong features in the cosmos. Fig.\\ \\ref{fighex} shows a predictive map of the hot and cold fluctuations in a $2\\pi/3$-twisted hexagonal prism. We have set the length of the fundamental domain to be ten times smaller in the twisted direction than along the face of the hexagon. The average large angle power in fluctuations is actually consistent with the data, although clearly this anisotropic space does not look like the sky we observe. A better statistic to discern patterns and correlations is badly needed \\cite{{lss},{lssb}}. The promising suggestions of \\cite{{bps},{css},{lssb}} may be the key and are discussed more in \\S \\ref{comph}. \\begin{figure} \\centerline{{\\psfig{file=tile2.eps,width=1.5in}}} \\centerline{{\\psfig{file=hexmodes.eps,width=1.75in}}} \\caption{Top: A guess at one mode. Bottom: A contour plot of the temperature fluctuation for a similar mode. \\label{onemode}} \\end{figure} We could have predicted certain features of the map of Fig.\\ \\ref{fighex}, even if we had not known the eigenmodes explicitly. A 2D slice through the 3D tiling of space is represented in Fig.\\ \\ref{onemode}. If we draw bands connecting opposite sides of the hexagons and highlight any overlaps, we can predict the imprint of one mode as shown on the top of Fig. \\ref{onemode}. Given that we do know the eigenmodes, we can show the actual contour plot of the hot and cold fluctuations for a similar mode on the bottom of Fig.\\ \\ref{onemode}. Comparing the guess with the actual contours shows our guess did quite well. The hexagonal shape of the universe is clearly seen. In actuality, there are many modes competing to imprint a pattern on the sky which blurs the signature hexagons. In Fig.\\ \\ref{mode2}, is another contour plot of hot and cold spots for a different mode which exhibits the $2\\pi/3$ twist through the prism. \\begin{figure} \\centerline{{\\psfig{file=hexmodes2.eps,width=1.75in}}} \\caption{A contour plot of the temperature fluctuation for a mode that winds through the twisted prism. \\label{mode2}} \\end{figure} The competition between fluctuations obscurs some features while enhancing others. The surface of last scatter cuts a sphere out of the full 3D space, an elliptic projection of which is given in the map of Fig. \\ref{fighex}. Can you see hexagons in the map? Almost. Is the $2\\pi/3$ twist in the space visible? We are currently developing ways of looking at the sky that pull the underlying patterns out of the noise \\cite{lssb}. ", "conclusions": "" }, "9803/astro-ph9803311_arXiv.txt": { "abstract": "We present the results of deep spectropolarimetry of two powerful radio galaxies at $z\\sim2.5$ (4C 00.54 and 4C 23.56) obtained with the W.M. Keck II 10m telescope, aimed at studying the relative contribution of the stellar and non-stellar components to the ultraviolet continuum. Both galaxies show strong linear polarization of the continuum between rest-frame $\\sim$1300-2000~\\AA, and the orientation of the electric vector is perpendicular to the main axis of the UV continuum. In this sense, our objects are like most 3C radio galaxies at $z\\sim1$. The total flux spectra of 4C 00.54 and 4C 23.56 do not show the strong P-Cygni absorption features or the photospheric absorption lines expected when the UV continuum is dominated by young and massive stars. The only features detected can be ascribed to interstellar absorptions by SiII, CII and OI. Our results are similar to those for 3C radio galaxies at lower $z$, suggesting that the UV continuum of powerful radio galaxies at $z\\sim2.5$ is still dominated by non-stellar radiation, and that young massive stars do not contribute more than $\\approx$50\\% to the total continuum flux at 1500~\\AA. ", "introduction": "High-$z$ radio galaxies (H$z$RGs) are observable to very high redshifts and can be used to study the formation and evolution of massive elliptical galaxies (see McCarthy 1993 for a review). One of the most controversial issues is the physical cause of the alignment between the radio source and UV continuum axes of the H$z$RGs (the so called `alignment effect', Chambers, Miley \\& van Breugel 1987, McCarthy et al. 1987). Two main competing scenarios have been proposed. The first is star formation induced by the propagation of the radio source through the ambient gas (see McCarthy 1993 and references therein); the second explains the alignment effect as the result of a hidden quasar whose radiation is emitted anisotropically and scattered towards the observer, producing strong linear polarization perpendicular to the radio-UV axis (Tadhunter et al. 1988; di Serego Alighieri et al. 1989). The latter scenario is closely related to the unification of powerful radio-loud AGN, and provides a way of testing it directly (see Antonucci 1993 and references therein). After the first detections of strong UV polarization in H$z$RGs obtained with 4m-class telescopes (di Serego Alighieri et al. 1989; Jannuzi \\& Elston 1991; Tadhunter et al. 1992; Cimatti et al. 1993), recent observations made with the Keck I 10m telescope have demonstrated the presence of spatially extended UV continuum polarization and of hidden quasar nuclei in some of the 3C radio galaxies at $0.72$. We have started a program of observations of these galaxies using spectropolarimetry at the Keck II 10m telescope, and in this Letter we report on the first two objects we have studied, concentrating on their continuum and absorption line properties. Throughout this paper we assume $H_0=50$ kms$^{-1}$ Mpc$^{-1}$ and $q_0=0$. ", "conclusions": "Our observations suggest that the UV spectra of 4C 23.56 and 4C 00.54 are not dominated by young massive stars, whereas the strong perpendicular polarization indicates the presence of a relevant scattered continuum, making 4C 23.56 and 4C 00.54 similar to the polarized 3C radio galaxies at 0.7$3$ have a major episode of star formation, and their AGN scattered component is diluted by the stellar light, but it becomes observable at lower $z$ when the starburst ceases. However, given the rapid evolution of the UV light from a starburst, it is also possible that 4C 41.17 simply represents a case dominated by the starburst rather than an evolutionary sequence. Future observation of a complete sample of H$z$RGs will help us to understand the nature of the alignment effect and the evolution of the host galaxies of powerful radio sources." }, "9803/astro-ph9803127_arXiv.txt": { "abstract": "We present high-quality HST/GHRS spectra in the Hydrogen L$\\alpha$ spectral region of Vega and Sirius-A. Thanks to the signal-to-noise ratio achieved in these observations and to the similarity of the two spectra, we found clear evidence of emission features in the low flux region, $\\lambda\\lambda$1190-1222\\,\\AA. These emission lines can be attributed unambiguously to \\ion{Fe}{ii} and \\ion{Cr}{ii} transitions. In this spectral range, silicon lines are observed in absorption. We built a series of non-LTE model atmospheres with different, prescribed temperature stratification in the upper atmosphere and treating \\ion{Fe}{ii} with various degrees of sophistication in non-LTE. Emission lines are produced by the combined effect of the Schuster mechanism and radiative interlocking, and can be explained without the presence of a chromosphere. Silicon absorption lines and the L$\\alpha$ profile set constraints on the presence of a chromosphere, excluding a strong temperature rise in layers deeper than $\\tau_{\\rm R} \\approx 10^{-4}$. ", "introduction": "On the main sequence, A-type stars are at a juncture point between hot and cool stars. While hot, massive stars undergo strong mass loss in fast winds ($\\dot{M}\\ge 10^{-9}$\\,M$_\\odot$/yr), cool stars show chromospheric activity connected to their subsurface convective layers. Both phenomena apparently disappear or become much weaker at spectral type A. Many studies have thus been devoted to the outer layers of A-type stars to search for indications of a wind or of stellar activity. Several attempts to detect signatures of weak winds in main-sequence A stars have been unsuccessful (e.g. Lanz~\\& Catala \\cite{lanz92}). Recently, however, a quite weak, blue-shifted absorption was detected in the \\ion{Mg}{ii} resonance lines of Sirius, and interpreted as a wind signature (Bertin et~al. \\cite{sirius3}). A mass loss rate of $\\dot{M}\\approx 10^{-12}$\\,M$_\\odot$/yr was derived, consistent with the idea that A-type star winds are radiatively-driven like the winds of hotter stars. On the cool side, a limit to chromospheric activity has been set at A7 (B\\\"ohm-Vitense~\\& Dettmann \\cite{bohmvitense80}, Marilli et~al. \\cite{marilli97}, Simon~\\& Landsman \\cite{simon97}). The most common diagnostics of chromospheres and winds are emission features. Therefore, we are not expecting emission lines in A stars, except cases where such lines arise from the circumstellar environment. High-quality ultraviolet spectra have become available with the {\\em Goddard High Resolution Spectrograph} (GHRS) aboard the {\\em Hubble Space Telescope} (HST). Even the core of strong resonance lines, including \\ion{H}{i} L$\\alpha$, can be observed with a reasonably good signal-to-noise ratio. This makes it possible to investigate in greater detail the line profile of strong resonance lines. They are the best tool to probe the outer layers of stars, being indeed formed very high in the atmosphere. In this respect, L$\\alpha$ is most interesting because it spans the largest range of depth of formation, from the far wing to the line core. This large variation in opacity also affects the formation of lines of other elements, especially close to L$\\alpha$ core. Such lines see a much lower local pseudo-continuum than lines outside L$\\alpha$, and will be formed much higher in the atmosphere than weak lines in other regions. \\begin{table*} \\caption[]{Observation log.} \\label{TabObs} \\begin{tabular}{lccccr} \\hline &&&&& \\\\ [-3mm] Target & Spectral Range & Date of Observation & GHRS Grating & Aperture & Exposure time \\\\ \\hline &&&&& \\\\ [-3mm] Sirius-A & 1188 \\AA\\ -- 1218 \\AA & 1996 Nov 20 & G140M & SSA & 1632.0 s \\\\ Sirius-A & 1278 \\AA\\ -- 1307 \\AA & 1996 Nov 20 & G140M & SSA & 217.6 s \\\\ Sirius-A & 1308 \\AA\\ -- 1337 \\AA & 1996 Nov 20 & G140M & SSA & 217.6 s \\\\ Vega & 1185 \\AA\\ -- 1222 \\AA & 1996 Dec 23 & G160M & SSA & 435.2 s \\\\ Vega & 1274 \\AA\\ -- 1311 \\AA & 1996 Dec 23 & G160M & SSA & 108.8 s \\\\ Vega & 1303 \\AA\\ -- 1341 \\AA & 1996 Dec 23 & G160M & SSA & 108.8 s \\\\ \\hline \\end{tabular} \\end{table*} In Sect. 2 and 3, we will describe our GHRS observations around L$\\alpha$ of two bright A stars, Vega and Sirius-A. We will in particular point out the presence of \\ion{Fe}{ii} and \\ion{Cr}{ii} emission lines between 1190 and 1222\\,\\AA. Bertin et~al. (\\cite{sirius2}) noticed the presence of emission features around L$\\alpha$ in a Cycle~1 GHRS spectrum of Sirius-A, originally recorded to derive the D/H abundance ratio in the local interstellar medium. This prompted us to repeat and extend these observations to investigate their origin. An explanation of these emission features is given in the second half of the paper. In Sect.~4, we describe our new non-LTE model atmospheres. We explore and set limits on a chromosphere (Sect.~5), and investigate non-LTE effects in \\ion{Fe}{ii} line formation (Sect.~6). ", "conclusions": "We have reported emission features in the L$\\alpha$ profile of Vega and Sirius-A. These emission lines have been attributed to \\ion{Fe}{ii} and \\ion{Cr}{ii} transitions. The identification appears quite secure because all the lines of several multiplets appear in emission. We have built non-LTE model atmospheres with different assumed temperature structures in the outer layers and incorporating \\ion{Fe}{ii} with different degrees of sophistication. We found that the emission features cannot be explained by a chromospheric temperature rise. To produce the observed \\ion{Fe}{ii} emissions, the temperature would have to increase in relatively deep layers, turning other lines into emission (e.g. \\ion{Si}{ii}, \\ion{Si}{iii} lines). However, we cannot exclude a chromospheric rise in shallower layers ($\\tau_{\\rm R} \\le 10^{-4}$) based on our present observations, in agreement with earlier results (Freire-Ferrero et~al. \\cite{freire83}). Non-LTE \\ion{Fe}{ii} line formation calculations with different model atoms have demonstrated that some \\ion{Fe}{ii} lines can turn into emission in the wavelength range between 1190 and 1240\\,\\AA. We stress that emission lines are predicted {\\em only} in this very low flux, central region of L$\\alpha$. This results from the combined effect of the Schuster mechanism and radiative interlocking. Some highly-excited levels are overpopulated by transitions occurring in a high-flux region, and preferentially de-excite in this region near L$\\alpha$. This mechanism explains the similarity of Vega and Sirius spectra. Differences between the two stars can also be understood with this mechanism. The higher heavy-element content in Sirius' photosphere results in depressing the flux, in particular near the flux maximum. The efficiency of the pumping is thus reduced, yielding generally weaker emissions in Sirius than in Vega. The details depend on the exact wavelength of the pumping transitions. The flatter L$\\alpha$ profile in Vega is also a consequence of the different metallicity. Lyman continuum heating must be more efficient in Vega's case (less heavy-element line opacity) yielding a somewhat higher temperature in the outer layers and a higher flux in the central region of L$\\alpha$. We believe that the origin of the \\ion{Cr}{ii} emission lines may be explained by similar mechanisms. While we cannot rule out that radiative interlocking is also effective in \\ion{Cr}{ii}, the difference between Vega and Sirius points to the Schuster mechanism being the major cause of the emission in this case. The lines are thus stronger in Sirius due to the larger chromium abundance, and are simply too weak in Vega to stand out of the noise. Although our model atmosphere calculations provide an explanation to an unexpected observation of emission lines in the spectrum of early A-type stars, we did not achieve a good fit to the L$\\alpha$ profile at this stage. It seems however likely that the flux observed in the central region of L$\\alpha$ may be explained by increasing somewhat the fraction of non-coherent scattering in the PRD approximation that we have used. Matching these observations would require (at least) non-LTE line-blanketed model atmospheres, treatment of L$\\alpha$ in partial redistribution tuning the ratio between coherent and non-coherent scattering, and improved, non-hydrogenic \\ion{Fe}{ii} photoionization cross-sections. Such an approach is necessary to gain a deeper insight into the outer layers of Vega and Sirius, and this certainly deserves further study. Finally, we did not find emission lines very close to the L$\\alpha$ core, especially in the 0.5\\,\\AA\\ blueward of the central wavelength. This implies fortunately that we have so far no reason to question the previous results on the local interstellar cloud (Bertin et~al. \\cite{sirius2}), and on the wind absorption feature (Bertin et~al. \\cite{sirius3})." }, "9803/astro-ph9803037_arXiv.txt": { "abstract": "We have preliminary results on the parallelization of a Tree-Code for evaluating gravitational forces in N-body astrophysical systems. For our T3D CRAFT implementation, we have obtained an encouraging speed-up behavior, which reaches a value of 37 with 64 processor elements (PEs). According to the Amdahl'law, this means that about 99\\% of the code is actually parallelized. The speed-up tests regarded the evaluation of the forces among $N = 130,369$ particles distributed scaling the actual distribution of a sample of galaxies seen in the Northern sky hemisphere. Parallelization of the time integration of the trajectories, which has not yet been taken into account, is both easier to implement and not as fundamental. ", "introduction": "Super computers are allowing a rapid development of numerical simulations of large N--body systems in Astrophysics. These systems are generally composed by both collisionless matter (such as: stars, galaxies, ...) and collisional matter (i.e. gas). Both phases are usually characterized by being self--gravitating, that is the dynamics of the bodies (stars or fluid elements) is strongly influenced by the gravitational field produced by the bodies themselves. This {\\it self-influence} is what makes the evaluation of the long--range gravitational force the heaviest computational task to perform in a dynamical simulation. In fact, the number of terms which has to be considered in a direct and trivial evaluation of all the interactions between bodies grows like $N^2$, and since many astrophysically realistic simulations require very large $N$ (greater than $10^5$), such a direct numerical evaluation seems hard to face with presently available computers. To overcome this problem various approximate techniques to compute gravitational interactions have been proposed. Among them, the Tree--code algorithm proposed by Barnes \\& Hut\\footnote{ Barnes J., Hut P. ``A hierachical $O(N\\log N)$ force calculation algorithm''. {\\it Nature}, vol. 324, p. 446 (1986).} is now widely used in Astrophysics because it does not require any spatial fixed grid (like, for example, methods based on the solution of Poisson's equation). This makes it particularly suitable to follow very inhomogeneous and variable (in time) situations, typical of self-gravitating systems out of equilibrium. In fact its intrinsic capability to give a rapid evaluation of forces allows spending more CPU-time to follow fast dynamical evolution, in contrast to other higher accuracy methods that are more suitable for other physical situations, e.g. dynamics of polar fluids, where the Coulomb term is present. With the help of the parallelization of our codes, we intend to increase by one or two order of magnitude the number of particles we can use to represent physical systems, in respect to that generally adopted on serial computers ($\\sim 10^4$). In particular our first scientifical aim is the study of close encounters between massive black holes and globular clusters. These latter are systems formed by more than $10^5$ stars gravitationally bounded in a spherical peaked distribution. Such a problem is important in the effort to understand better the nature and formation mechanisms of the {\\it Active Galactic Nuclei}\\ \\footnote{ Capuzzo-Dolcetta R., Miocchi P., ``Galactic Nuclei Activity Sustained by Globular Cluster Mass Accretion'', {\\it PaSS} (1998) in press.}. We hope parallelization makes possible to represent each star with a single particle, in a one--to--one correspondence. This fact clearly will make simulations much more physically meaningful. ", "conclusions": "" }, "9803/astro-ph9803171_arXiv.txt": { "abstract": "The isothermal gravitational collapse and fragmentation of a molecular cloud region and the subsequent formation of a protostellar cluster is investigated numerically. The clump mass spectrum which forms during the fragmentation phase can be well approximated by a power law distribution $dN/dM \\propto M^{-1.5}$. In contrast, the mass spectrum of protostellar cores that form in the centers of Jeans unstable clumps and evolve through accretion and $N$-body interaction is best described by a log-normal distribution. Assuming a star formation efficiency of $\\sim\\!10\\;\\!$\\%, it is in excellent agreement with the IMF of multiple stellar systems. ", "introduction": "\\label{sec:intro} Understanding the processes leading to the formation of stars is one of the fundamental challenges in astronomy and astrophysics. However, theoretical models considerably lag behind the recent observational progress. The analytical description of the star formation process is restricted to the collapse of isolated, idealized objects (Whitworth \\& Summers 1985). Much the same applies to numerical studies (e.g.~Boss 1997, Burkert et al.~1997 and reference therein). Previous numerical models that treated cloud fragmentation on scales larger than single, isolated clumps were strongly constrained by numerical resolution. Larson (1978), for example, used just 150 particles in an SPH-like simulation. Whitworth et al.~(1995) were the first who addressed star formation in an entire cloud region using high-resolution numerical models. However, they studied a different problem: fragmentation and star formation in the shocked interface of colliding molecular clumps. While clump-clump interactions are expected to be abundant in molecular clouds, the rapid formation of a whole star cluster requires gravitational collapse on a size scale which contains many clumps and dense filaments. Here, we present a high-resolution numerical model describing the dynamical evolution of an entire {\\em region} embedded in the interior of a molecular cloud. We follow the fragmentation into dense protostellar cores which form a hierarchically structured cluster. ", "conclusions": "\\label{sec:summary} Large-scale collapse and fragmentation in molecular clouds leads to a hierarchical cluster of condensed objects whose further dynamical evolution is extremely complex. The agreement between the numerically-calculated mass function and the observations strongly suggests that gravitational fragmentation and accretion processes dominate the origin of stellar masses. The final mass distribution of protostellar cores in isothermal models is a consequence of the chaotic kinematical evolution during the accretion phase. Our simulations give evidence, that the star formation process can best be understood in the frame work of a probabilistic theory. A sequence of statistical events may naturally lead to a log-normal IMF (see e.g.~Zinnecker 1984, Adams \\& Fatuzzo 1996; also Price \\& Podsiadlowski 1995, Murray \\& Lin 1996, Elmegreen 1997)." }, "9803/chao-dyn9803019_arXiv.txt": { "abstract": "We study the dynamical and statistical behavior of the Hamiltonian Mean Field (HMF) model in order to investigate the relation between microscopic chaos and phase transitions. HMF is a simple toy model of $N$ fully-coupled rotators which shows a second order phase transition. The canonical thermodynamical solution is briefly recalled and its predictions are tested numerically at finite $N$. The Vlasov stationary solution is shown to give the same consistency equation of the canonical solution and its predictions for rotator angle and momenta distribution functions agree very well with numerical simulations. A link is established between the behavior of the maximal Lyapunov exponent and that of thermodynamical fluctuations, expressed by kinetic energy fluctuations or specific heat. The extensivity of chaos in the $N \\to \\infty$ limit is tested through the scaling properties of Lyapunov spectra and of the Kolmogorov-Sinai entropy. Chaotic dynamics provides the mixing property in phase space necessary for obtaining equilibration; however, the relaxation time to equilibrium grows with $N$, at least near the critical point. Our results constitute an interesting bridge between Hamiltonian chaos in many degrees of freedom systems and equilibrium thermodynamics. ", "introduction": "Many-particle systems can show collective behavior when the average kinetic energy is small enough. This collective macroscopic behavior can coexist with chaos at the microscopic level. Such a behavior is particularly evident for systems that have a phase transition, for which a nonvanishing order parameter measures the degree of macroscopic organization, while at the microscopic level chaotic motion is a source of disorder. The latter can induce non trivial time dependence in the macroscopic quantities, and it would be desirable to relate the time behavior of such quantities and their fluctuations to the chaotic properties of microscopic motion, measured through the Lyapunov spectrum. A naive idea is that an increase of chaos as the energy (temperature) is increased should be accompanied with a growth of fluctuations of some macroscopic quantity. These should be maximal at the critical point and then drop again at high energy. In this paper we study a model of $N$ fully-coupled Hamiltonian rotators which realizes such a behavior, it has been called Hamiltonian Mean Field (HMF) model~\\cite{antoni,latora}. It can also be considered as a system of interacting particles moving on a circle. This system has a second order phase transition and in the ordered phase the rotators are clustered; the high temperature phase is a gaseous one, with the particles uniformly distributed on the circle. It has been shown in ref.~\\cite{latora} that the maximal Lyapunov exponent grows up to the critical energy density $U_c$ and then drop to zero in the whole high temperature phase in the $N \\to \\infty$ limit. Correspondingly one observes a growth of kinetic energy fluctuations up to the critical point and then a phase of vanishing fluctuations. Finite $N$ effects complicate this simple picture. In the high temperature phase the maximal Lyapunov exponent vanishes quite slowly (with $N^{-1/3}$) and finite size effects influence the first region below the critical point. In this region the system displays metastability: starting far from equilibrium, this is reached in a time $\\tau_r$ which grows with $N$. On the contrary, the extremely low energy phase is characterized by a weak $N$ dependence, with the maximal Lyapunov exponent $\\lambda_1$ which behaves as $\\lambda_1 \\sim \\sqrt{U}$. Although the model is extremely simplified, it shares many features with more complex models, for which the relation between chaotic motion at the microscopic level and collective macroscopic properties has been studied. Let us mention studies in solid state physics and lattice field theory~\\cite{solid,nayak,dellago,yama,lapo}. However, it has been actually in nuclear physics~\\cite{ata,cmd}, where there is presently a lively debate on multifragmentation phase transition~\\cite{ata,cmd,gsi,eos,bondgro,bond,perco,cmd1,mastinu}, that the interest in the connection between chaos and phase transitions has been revived. In this case in fact, an energy/temperature relation quite close to the HMF model has been observed~\\cite{gsi} and critical exponents have been measured experimentally~\\cite{eos}. Statistical thermodynamical models~\\cite{bondgro} and percolation approaches ~\\cite{perco} have been proved to give a good description of the experimental data, though the dynamics is missing. On the other hand classical molecular dynamics models \\cite{cmd,bond,cmd1} seem to contain all the main ingredients, but have the disadvantage that a detailed understanding of the dynamics can be too complicated. In this respect, the HMF model can be very useful in clarifying some general dynamical features which could be eventually compared with real experimental data. In fact, when studying nuclear multifragmentation, one deals with excited clusters of 100-200 particles interacting via long-range (nuclear and Coulomb) forces. Quantum effects are relevant only at very low energy. In fact in the nuclear case, at very low energy, $T$ is not linear in $U$, but grows as $\\sqrt{U}$ because nucleons are fermions~\\cite{gsi,bondgro}. However, a classical picture should be quite realistic in the critical region where the excitation energy is substantial ~\\cite{mastinu}. In this paper we present new numerical data concerning both statistical quantities, like specific heat and distribution functions, and chaotic probes, like Lyapunov spectra and Kolmogorov-Sinai entropy. Moreover, we add to the theoretical analysis of the model a thorough treatment of differences in the fluctuating quantities between the canonical and microcanonical ensembles. We also investigate in detail the relaxation to equilibrium and compare numerical results with a complete self-consistent Vlasov calculation of distribution functions. Finally a comparison of numerically obtained maximal Lyapunov exponents with theoretical formulas is attempted. The paper is organized as follows. In Sec. 2 we briefly discuss the details of the HMF model. The equilibrium statistical mechanics and the continuum Vlasov solution are described in Sects. 3 and 4 respectively. In Sec. 5 we discuss the relaxation to equilibrium and in Sec. 6 we present the numerical calculations of the Lyapunov spectra and Kolmogorov-Sinai entropy as a function of the energy and $N$. Analytical estimates are discussed in Sec. 7 and conclusions are drawn in Sec. 8. ", "conclusions": "We have investigated the dynamical and statistical behavior of a system with long-range forces showing a second order phase transition. Both the maximal Lyapunov exponent $\\lambda_1$ and the Kolmogorov-Sinai entropy density $S_{KS}/N$ are peaked at the phase transition point, where kinetic energy fluctuations and specific heat are maximal. There is actually a small shift to lower energies due to finite size effects. The latter are present also in the Lyapunov spectra and in the Kolmogorov-Sinai entropy. Above the phase transition point, both $\\lambda_1$ and $S_{KS}$ vanish as $N \\to \\infty$. We think that this toy model contains some important ingredients to understand the behavior of macroscopic order parameters when dynamical chaos is present at the microscopic level. Most of our findings are probably common to other Hamiltonian systems showing second order phase transitions. In particular our results could be very important in order to understand the relaxation to the equilibrium solution and the success of statistical approaches in describing the nuclear multifragmentation phase transition. \\begin{ack} We thank M.C. Firpo for communicating us her results before publication and P. Holdsworth for interesting suggestions. We thank A. Torcini for many useful discussions and a careful reading of the text. A.R. thanks the Centre for Theoretical Physics of MIT for the kind hospitality and M. Robnik for stimulating discussions during his visits at CAMTP in Maribor, Slovenia. V.L. and S.R. thank INFN for financial support. S.R. thanks CIC, Cuernavaca, Mexico for financial support. This work is also part of the European contract No. ERBCHRXCT940460 on ``Stability and universality in classical mechanics\". {\\it In the 60's, Boris Chirikov was also an explorer of the (no man's land at that time) relation between chaotic motion and statistical behavior in classical systems with many degrees of freedom. We hope that he will be interested by this work.} \\end{ack}" }, "9803/astro-ph9803084_arXiv.txt": { "abstract": "Using direct N-body simulations which include both the evolution of single stars and the tidal field of the parent galaxy, we study the dynamical evolution of globular clusters and rich open clusters. We compare our results with other N-body simulations and Fokker-Planck calculations. Our simulations, performed on the GRAPE-4, employ up to 32,768 stars. The results are not in agreement with Fokker-Planck models, in the sense that the lifetimes of stellar systems derived using the latter are an order of magnitude smaller than those obtained in our simulations. For our standard run, Fokker-Plank calculations obtained a lifetime of 0.28 Gyr, while our equivalent $N$-body calculations find $\\sim4$ Gyr. The principal reason for the discrepancy is that a basic assumption of the Fokker-Plank approach is not valid for typical cluster parameters. The stellar evolution timescale is comparable to the dynamical timescale, and therefore the assumption of dynamical equilibrium leads to an overestimate of the dynamical effects of mass loss. Our results suggest that the region in parameter space for which Fokker-Planck studies of globular cluster evolution, including the effects of both stellar evolution and the galactic tidal field, are valid is limited. The discrepancy is largest for clusters with short lifetimes. ", "introduction": "Theoretical models for the evolution of star clusters are generally too idealized for comparison with observations. However, detailed model calculations with direct $N$-body methods are not feasible for real globular clusters, even with fast special-purpose computers such as GRAPE-4 (Makino et al.\\ 1997)\\nocite{1997ApJ...480..432M} or advanced parallel computers (Spurzem \\& Aarseth 1996).\\nocite{1996MNRAS.282...19S} If we could scale the results of $N$-body simulations with relatively small numbers of particles (such as $\\sim30,000$) to real globular clusters, then it would become feasible to perform computations with relatively small numbers of particles and still derive useful qualitative conclusions about larger, more massive systems. However, to determine the proper scaling is difficult because the ratio between two fundamental time scales, the relaxation times and the dynamical time, is proportional to $N$. In typical globular clusters, this ratio exceeds $10^3$ and the two time scales are well separated. In $N$-body simulations, the ratio is generally much smaller. The inclusion of realistic effects such as mass loss due to stellar evolution and the effect of galactic tidal fields (with the galaxy approximated as a point mass, but also with the inclusion of disc shocking) further complicate the scaling problem (see, e.g., Heggie 1996).\\nocite{heg96} A proper treatment of stellar evolution is particularly problematic, since its characteristic timescale changes as stars evolve. Chernoff \\& Weinberg (1990, CW90)\\nocite{cw90} performed an extensive study of the survival of star clusters using Fokker-Planck calculations which included 2-body relaxation and some rudimentary form of mass loss from the evolving stellar population. In their simulations the number of particles is not specified. Their models are defined by the initial half-mass relaxation time and by the initial mass function of the cluster. Since their models do not specify the number of stars per cluster, each of their model calculations corresponds to a one-dimensional series of models, when plotted in a plane of observational values, such as total mass versus distance to the galactic center (Fig.\\ 1). All points of the solid line in that figure correspond to a single calculations by CW90, since they have an identical relaxation time. As we will see later, it is useful to consider other series of models, for which the crossing time is held constant while varying the mass. An example of such a series is indicated by the dashed line in Fig.\\ 1. The shapes of these lines are derived under the assumption of a flat rotation curve for the parent galaxy. The main conclusion of CW90 was that the majority of the simulated star clusters dissolve in the tidal field of the galaxy within a few hundred million years. Fukushige \\& Heggie (1995, FH95)\\nocite{fh95} studied the evolution of globular clusters using direct $N$-body simulation, using the same stellar evolution model as used by CW90. They used a maximum of 16k particles and a scaling in which the dynamical timescale of the simulated cluster was the same as that of a typical globular cluster, corresponding to one of vertical lines in Fig. 1. FH95 found lifetimes much longer than those in CW90's Fokker-Planck calculations, for the majority of the models used in CW90. However, the reason for the discrepancy is rather unclear, because the calculations of FH95 and those of CW90 differ in several important respects. The relaxation times differ because FH95 held the cluster crossing time fixed in scaling from the model to the real system. However, the crossing times themselves are also different, since the crossing time is by definition zero in a Fokker-Planck calculation. Finally, the implementation of the galactic tidal field is also quite different. CW90 used a simple boundary condition in energy space (spherically symmetric in physical space), in which stars were removed once they acquired positive energy, but the underlying equations of motion included no tidal term. FH95 adopted a much more physically correct treatment, including tidal acceleration terms in the stellar equations of motion and a proper treatment of centrifugal and coriolis forces in the cluster's rotating frame of reference (see FH95). \\begin{figure} \\centerline{ \\psfig{file=fig_isomodels.ps,bbllx=570pt,bblly=40pt,bburx=110pt,bbury=690pt,height=5cm,angle=-90}} \\caption {Cluster mass versus the distance to the galactic center. The solid line indicates the model parameters for which the relaxation time is constant (iso relaxation time); the dashed line indicates the initial conditions for which the crossing time of the star cluster is constant (iso crossing time) } \\label{isomodels}\\end{figure} In order to study the behavior of star clusters with limited numbers of stars, and to compare with the results of the Fokker-Planck simulations of CW90, we selected one of their models and perform a series of collisional N-body simulations in which the evolution of the individual stars is taken into account. According to CW90 the results should not depend on the number of stars in the simulation as long as the relaxation time is taken to be the same for all models. It is, among others, this statement which we intend to study. We find that for this set of initial conditions Fokker-Planck models do not provide a qualitatively correct picture of the evolution of star clusters. The effects of the finite dynamical time scale are large, even for models whose lifetime is several hundred times longer than the dynamical time. The main purpose of this paper is to study the survival probabilities of star clusters containing up to a few tens of thousands of single stars, in order to gain a deeper understanding of the influence of the galactic tidal field and the fundamental scaling of small $N$ clusters to larger systems. Only single stars are followed; primordial binaries are not included. The computation of gravitational forces is performed using the GRAPE-4 (GRAvity PipE, see \\cite{emf+93}, a special-purpose computer for the integration of large collisional $N$-body systems). Hardware limitations (speed as well as storage) restrict our studies to $\\aplt 32$k particles. The dynamical model, stellar evolution, initial conditions and scaling are discussed in Sect. 2. Section 3 reviews the software environment and the GRAPE-4 hardware, and discusses the numerical methods used. The results are presented in Sect. 4 and discussed in Sect. 5; Sect. 6 sums up. ", "conclusions": "We have followed the evolution of a star cluster, to the point of dissolution in the tidal field of the parent galaxy, taking into account both the effects of stellar dynamics and of stellar evolution. Our calculations are based on direct $N$-body integration, coupled to approximate treatments of stellar evolution. Our results differ greatly from those obtained with Fokker-Planck calculations, as presented by CW90: their model clusters dissolve after a few times $10^8$ years, whereas our equivalent model clusters live at least ten times longer. As we discussed in the previous section, a number of different reason conspire to produce such a drastic difference. Our hope was that we would be able to find a way to bridge our $N$-body results and previous results based on Fokker-Planck approximations. The fact that the GRAPE-4 special-purpose hardware allowed us to model much larger numbers of particles, reaching to within an order of magnitude of that of real globular clusters, seemed to indicate that it would finally be possible to make a firm connection between the two types of simulations. However, our results indicate that no clear process of extrapolation has emerged yet. Even within the different runs we have studied, extrapolation from the smaller to the larger number of stars would have resulted in rather large errors. This suggests that further extrapolation will suffer from the same fate. In the present paper, we have studied in detail a single model. However, the way the result depends on the time scaling might be different for other models. In a subsequent paper, we plan to carry out a systematic study, similar to the one we have presented here, for a much wider range of initial models." }, "9803/hep-ph9803270_arXiv.txt": { "abstract": "The idea that the universe might be open is an old one, and the possibility of having an open universe arise form inflation is not new either. However, a concrete realization of a consistent single-bubble open inflation model is known only recently. There has been great progress in the last two years in the development of models of inflation consistent with observations in such an open universe. In this overview I will describe the basic features and the phenomenological consequences of such models, making emphasis in the predictions of the CMB temperature anisotropies that differ from ordinary inflation. ", "introduction": "The idea that the universe might be open is an old one, see e.g. \\cite{Peebles}. Early attempts to accomodate standard inflation in an open universe~\\cite{ratra} failed to realize that in usual inflation homogeneity implies flatness~\\cite{turner}, due to the Grishchuck-Zel'dovich effect~\\cite{GZE}. The possibility of having a truly open universe arise form inflation is not new either, see \\cite{Gott}, via the nucleation of a single bubble in de Sitter space. However, a concrete realization of a consistent model is known only recently, the single-bubble open inflation model~\\cite{singlebubble,LM}. Soon afterwards there was great progress in determining the precise primordial spectra of perturbations~[8-19], most of it based on quantum field theory in spatially open spaces. Simultaneously there has been a large effort in model building~\\cite{LM,Green,induced,GBL} and constraining the existing models from observations of the temperature power spectrum of cosmic microwave background (CMB) anisotropies~[21-24]. In this review talk I will concentrate in model building and constraints from CMB anisotropies. We will describe the nature of the various primordial perturbations and give the corresponding spectra, without deriving them from quantum field theoretical arguments. The interested reader should find this in the literature. We will then compute the corresponding angular power spectra of temperature anisotropies in the CMB. Furthermore, we will give a review of the different single- and mutiple-field open inflation models and constrain their parameters from present observations of the CMB anisotropies. Sometimes this is enough to rule out some of the models. Finally, we will describe how future observations of the CMB temperature and polarization anisotropies might be able to decide among different inflationary models, both flat and open inflation ones. ", "conclusions": "Single-bubble open inflation is an ingenious way of reconciling an infinite open universe with the inflationary paradigm. In this scenario, a symmetric bubble nucleates in de Sitter space and its interior undergoes a second stage of slow-roll inflation to almost flatness. In the near future, observations of the microwave background with the new generation of satellites, MAP and Planck, will determine with better than 1\\% accuracy whether we live in an open universe or not. It is therefore crucial to know whether inflation can be made compatible with such a universe. Single-bubble open inflation models provide a natural scenario for understanding the large scale homogeneity and isotropy. Furthermore, these inflationary models generically predict a nearly scale invariant spectrum of density and gravitational wave perturbations, which could be responsible for the observed CMB temperature anisotropies. Future observations could then determine whether these models are compatible with the observed features of the CMB power spectrum. For that purpose it is necessary to know the predicted power spectrum with great accuracy. Open models have a more complicated primordial spectrum of perturbations, with extra discrete modes and possibly large tensor anisotropies. In order to constrain those models we have to compute the full spectrum for a large range of parameters. In this review we have shown that the simplest single-field models of open inflation are not only fine tuned, but actually ruled out because they induce too large tensor anisotropies in the CMB, which is incompatible with present observations. On the other hand, two-field models generically do not lead to infinite open universes, as previously thought, but to an ensemble of very large but finite inflating `islands'. Each one of these islands will be a quasi-open universe. We may happen to live in one of those patches, where the universe {\\em appears} to be open. This new effect, semiclassical in origin, was recently discussed in Ref.~\\cite{quasi} where it was found that many of the present models are in fact quasi-open. This does not mean that they are not good cosmological models. If the co-moving size of the inflating islands is sufficiently large, then the resulting semiclassical anisotropy may be unobservable. We have shown however that such a component imposes very stringent constraints on the models. Most of them have a narrow range of parameters for which they are compatible with observations. It is perhaps worth mentioning here some alternative proposals (not single-bubble) for the generation of an open universe in the context of inflation. First of all, the group of Roma~\\cite{Roma} proposed a model based on higher order gravity that induces bubble nucleation and later percolation, resulting in a distribution peaked at $\\Omega_0\\simeq0.2$. Perhaps the most striking recent results are those of Hawking-Turok~\\cite{HawTur} and Linde~\\cite{creation}, who claim that an open inflationary universe could have been created directly from the vacuum, without the intermediate de Sitter phase." }, "9803/astro-ph9803059_arXiv.txt": { "abstract": "Numerical simulations of galaxy clusters including two species -- baryonic gas and dark matter particles -- are presented. Cold Dark Matter spectrum, Gaussian statistics and flat universe are assumed. The dark matter component is evolved numerically by means of a standard {\\it particle mesh} method. The evolution of the baryonic component has been studied numerically by using a multidimensional (3D) hydrodynamical code based on {\\it modern high resolution shock capturing} techniques. These techniques are specially designed for treating accurately complex flows in which shocks appear and interact. With this picture, the role of shock waves in the formation and evolution of rich galaxy clusters is analyzed. Our results display two well differenced morphologies of the shocked baryonic matter: filamentary at early epochs and quasi-spherical at low redshifts. ", "introduction": "Galaxy clusters are the largest systems gravitationally bounded in the Universe. Their study has been a fashion topic in Cosmology since last years. Work on topics related with galaxy clusters is worthly to: i) understand the formation, evolution, dynamics and morphology of these systems, ii) learn on the physical processes involved in them, and iii) find out some information concerning with fundamental parameters in Cosmology as density parameter ($\\Omega$) , Hubble's constant ($H$), and the spectrum of the primordial density field. During last years technical improvements have produced huge quantity of data about galaxy clusters. Let us mention the new galaxy surveys (Guzzo 1996, and references therein) and the extensive observation in X-rays using satellites as ROSAT or ASCA. These huge volume of data strongly motivates a lot of theoretical work trying to explain the observational results. From the theoretical point of view, numerical simulations are the best tools to understand physics involved in galaxy clusters. At the beginning, numerical simulations of galaxy clusters where performed using N-body techniques. Since then, they have been extensively used and have produced important results ( see, e.g, Efstathiou et al. 1985, Bertschinger \\& Gelb 1991, Xu 1995). Next step in the full description of galaxy clusters was to introduce in the picture a baryonic component. The numerical methods developed in order to deal with baryonic matter were more sofisticated and expensive in computational resources. As a consequence, it was not possible to carry out numerical simulations with two species (dark matter and baryonic gas) until late eighties. Cosmological hydrodynamic codes have been usually classified in two main categories: a) the so-called Lagrangian methods, like the {\\it Smoothed Particles Hydrodynamics} (SPH) or ulterior extensions based on them, and b) Eulerian codes. SPH methods were first proposed by Gingold \\& Monaghan (1977), and Lucy (1977). Among the best features of this technique, it should be pointed out its high resolution in dense regions. This property is directly derived from its Lagrangian character. The first implementations of SPH techniques had some weak points: i) The low density regions were badly described due to the Lagrangian character of the method. ii) Discontinuities and strong gradients were poorly solved and an important diffusion was introduced. iii) They were not conservative. Nevertheless, these previous problems were overcome in the modern implementations of these techniques. Improved SPH techniques have been widely developed for cosmological applications (see, e.g., Evrard 1988, Hernquist \\& Katz 1989, Navarro \\& White 1993, Gnedin 1995). Numerical cosmological codes using an Eulerian approach to study baryonic gas inside galaxy clusters have been also developed. Some of these hydro-codes use {\\it artificial viscosity} in order to deal with shock waves (Cen 1992, Anninos et al. 1994). These techniques require a good calibration of the free parameters which are introduced by hand and state some numerical problems. Recently, a new family of finite difference methods, which use Eulerian approaches and avoid artificial viscosity, has been developed in numerical Cosmology. They are the so-called {\\it high resolution shock capturing methods} (HRSC), the modern extensions of the original Godunov's idea (1959). According to the Riemann solver and the procedure in order to achieve spatial accuracy, we can distinguish three groups: 1) the ones following Harten's scheme (1983), like Ryu et al. (1993), 2) those using the analytical solution of the Riemann problem for the Newtonian dynamics of ideal gases and the PPM scheme described by Collela \\& Woodward (1984) , like Bryan et al. (1994), and 3) the codes using Roe's Riemann solver (Roe 1981) plus the MUSCL or PPM cell reconstruction, like in Quilis et al. (1996). In this last reference, the code used in present paper is described and tested appropriately. An exhaustive comparison among all these kinds of cosmological hydrodynamic codes can be found in Kang et al. (1994). Due to the Eulerian character of our code, it does not show --in dense regions-- a resolution as good as the Lagrangian ones, and it requires more computational resources. However, HRSC schemes --by construction-- have excellent properties in order to deal with shocks, discontinuities, and strong gradients. HRSC techniques typically solve shocks in two cells. Due to their intrinsic properties, the detection of shocks is independent on the number of cells used in the simulations. It should be pointed out that this last property is really important when three-dimensional simulations are carried out. In these simulations the size of the grid is a stringent constraint due to its high cost in computational resources. Moreover, these methods are conservative by construction, that is , quantities which should be physically conserved are numerically conserved up to the order of the method. It should be also noticed that these methods show good results in extreme low density regions (Einfeldt et al. 1991). As it has been pointed out by several authors, the role of shock waves can be extremely important in order to understand the heating processes in the intracluster medium (ICM). In this paper we are interested in understanding and quantifying the role of shocks. In order to do that it is crucial to use numerical codes able to manage with complex flows. Yes, indeed, one of the important features of HRSC techniques is just to treat numerically shocks and strong discontinuities giving sharp profiles (in a few numerical cells, as we have mentioned above) independently of the size of the grid. Hence, formation, evolution, and interaction of shocks in 3D flows can be analyzed accurately with HRSC schemes, and, consequently, their use is absolutely justified in order to study shocks and their consequences on the ICM's dynamics. Hereafter, $t$ stands for the cosmological time, $t_0$ is the age of the Universe, $a(t)$ is the scale factor of a flat background. Function $\\dot{a}/a$ is denoted by $H$, where the dot stands for the derivative of $a$ with respect to the cosmological time. Hubble constant is the present value of $H$; its value in units of $100 \\ Km \\ s^{-1} \\ Mpc^{-1}$ is the reduced Hubble constant $h$. In our computations we have assumed $h=0.5\\, $. Velocities are given in units of the speed of light. Baryonic, dark matter, and background mass density are denoted by $\\rho_{_b}$, $\\rho_{_{DM}}$, and $\\rho_{_{B}}$, respectively. The density contrast is $\\delta_b=(\\rho_b-\\rho_{_{B}})/\\rho_{_{B}}$ for baryonic matter, analogously is defined $\\delta_{_{DM}}$ for dark matter. The plan of this paper is as follows: In Section 2, our numerical cluster model is described. In Section 3, the results of the simulations are analyzed. Finally, a general discussion is presented in Section 4. ", "conclusions": "In this paper we have used some numerical techniques recently applied to Cosmology. These techniques , HRSC, seems to be the most suitable in order to study the role of shocks in galaxy cluster evolution. The choice is justified by their properties to handle shocks. The capability of these techniques to capture shocks with very small diffusion is independent of the resolution used in the numerical simulations. Hence, by construction, shocks are captured even using coarse grids. This property is crucial in 3D applications. Previous sections illustrate the fact that non adiabatic processes, due to shocks, take an important role in the description of the ICM. In the model presented in this paper, that is, a baryonic fluid plus dark matter component coupled gravitationally, shocks are able to heat the ICM until values compatible with observational data. The calculations have been carried out in two cases: with and without cooling processes. This procedure allows us to distinguish between non-adiabatic effects coming from shocks and the ones from cooling. The role of the cooling, even when it could be important in other scenarios, is irrelevant for the simulations considered in this paper, while shocks play the most important role. In the picture describing the dynamics of the baryonic component, there are some clues showing the presence of shocks. Examining the quantities sensitive to shocks, all of them evidence the formation of a quasi-spherical shock. This shock seems to arise around $z\\sim 2$ at the cluster center and moves outwards. Nevertheless, some irregular shocks could form at $z \\geq 2$. This conclusion is supported by the behaviour of the entropy profiles (see Fig. 5), and the existence of shocked cells at these times (see Fig. 8). The quasi-spherical shock would form from the collapse of the quasi-spherical global structure , while other smaller shocks -- with a filamentary morphology -- would arise from some collapsing substructure and merging processes. In short, previous discussion manifest two different regimes in the shock formation. It should be noticed that the structures simulated in this paper correspond, due to the initial conditions, to a large Abell cluster. For this kind of clusters, gravitational collapse is fast and the dynamics is violent. Shocks form earlier and are stronger than in others smaller cluster-like objects ($ < 3\\sigma$). Some discussion on the numerical resolution of the simulations is needed. The one used in present paper ($\\sim 0.3 Mpc$) is not enough to simulate the very center of the clusters and galaxy formation, but it suffices to study the role of the shocks in ICM. It should be kept in mind that HRSC techniques are able to resolve shocks even with coarse grids. Nevertheless, higher resolution would be desiderable to perform more complete simulations. Improvements in numerical resolution will introduce smaller scales in the problem, as a consequence, the physics of the model should be enriched in order to describe this new scenario. Chemical reactions and radiative transfer should be considered." }, "9803/astro-ph9803090_arXiv.txt": { "abstract": "We present a scenario for the formation and evolution of disk galaxies within the framework of an inflationary cold dark matter universe, and we compare the results with observations ranking from the present-day up to $z\\sim 1$. The main idea in this scenario is that galactic disks are built-up inside-out by gas infall with an accretion rate driven by the cosmological mass aggregation history (MAH). In Avila-Reese et al. (1997) the methods to generate the MAHs of spherical density fluctuations from a Gaussian random field, and to calculate the gravitational collapse and virialization of these fluctuations, were presented. Assuming detailed angular momentum conservation during the gas (5\\% of the total mass) contraction, a disk in centrifugal equilibrium is built-up within the forming dark matter halo. The primordial angular momentum is estimated through the Zel'dovich approximation and normalized to the spin parameter $% \\lambda $ given by analytical and numerical studies. The disk galactic evolution is followed through a physically self-consistent approach which considers (1) the gravitational interactions among the dark halo, the stellar and gas disks, and a bulge; (2) the turbulence and energy balance of the interstellar medium; (3) the star formation process due to gas disk gravitational instabilities; and (4) the secular formation of a bulge due to the gravitational instabilities of the stellar disk. We find that the main disk galaxy properties and their correlations are basically established by the combination of three fundamental physical factors: the mass, the MAH, and the spin parameter $\\lambda $. Models calculated for a statistically significant range of values for these factors predict nearly exponential disk surface brightness profiles with realistic central surface brightnesses $\\mu _{B_0},$ and scale lengths (including low surface brightness galaxies), nearly flat rotation curves, and negative gradients in the B-V color index radial distribution. The main trends across the Hubble sequence of the global intensive properties such as B-V, $\\mu _{B_0},$ the gas fraction $f_g$, and the bulge-to-disk ratio b/d, are reproduced. For a given mass (luminosity) B-V correlates with the maximum circular velocity, and this correlation is in agreement with the scatter of the Tully-Fisher relation. We interpret the observed color-magnitude, and ``color '' Tully-Fisher relations as a result of the empirical dependence of extinction on luminosity (mass). The model properties tend to form a biparametrical sequence, where B-V and $\\mu _{B_0}$ could be the two parameters. The star formation history depends on the MAH and on the $\\lambda$ parameter. A maximum in the star formation rate for most of the models is attained at $z\\sim 1.5-2.5$, where this rate is approximately 2.5-4.0 times larger than the present one. The scale radii and the bulge-to-total ratio decrease with $z$, while $\\mu _{B_0}$ increase. The B-band TF relation remains almost the same at different redshifts. Our scenario of disk galaxy formation and evolution reveals that the cosmological initial conditions are able to determine the main properties of disk galaxies across the Hubble sequence and predict evolutionary features for the present-day dominant galaxy population that are in agreement with very recent deep field observational studies ", "introduction": "The understanding of the formation and evolution of galaxies is one of the clearest challenges of contemporary astrophysics and cosmology. Since galaxies are both cosmological and astronomical objects, two general approaches can be used in order to study their formation and evolution (e.g., Renzini 1994): ({\\it i}) the deductive approach, through which, starting from some initial conditions given by a theory of cosmic structure formation, one tries to follow the evolutionary processes until the reconstruction of the observable properties of the galaxies and ({\\it ii}) the inductive approach, in which, starting from the present-day properties of galaxies, and through galactic evolutionary models, one tries to reconstruct the initial conditions of galaxy formation; the increasing observational data on galaxies at intermediate and high redshifts will enrich this approach with crucial constraints. Most of current theories about cosmic structure formation are based on the gravitational paradigm and on the inflationary cold dark matter (CDM) cosmological models. Since these models predict more power for the small density fluctuation scales than for the larger ones, cosmic structures build up hierarchically, through a continuous aggregation of mass. From the point of view of the galaxy cosmogony, a crucial question is whether this aggregation occurs through violent mergers of collapsed substructures and/or through a gentle process of mass aggregation. This question depends on the statistical distribution of the density fluctuation field and on its power spectrum. Nevertheless, even if the dark matter structures assemble through chaotic and violent mergers of subunits, the baryon gas, because of the reheating due to the shocks implied in the collapse, virialization and star formation (SF) feedback processes, will tend to aggregate around the density peaks in a more (spatially) uniform fashion than dark matter do it. Within the framework of the hierarchical clustering theory, from the most general point of view, two could be the galaxy formation scenarios. In one case, the main properties of galaxies, including those which define their morphological types, are supposed to be basically the result of a given sequence of mergers. This picture, that we shall call the{\\it \\ merger scenario,} has been widely applied in semianalytical models of galaxy formation where galaxies are constructed from the cosmological initial conditions through preconceived recipes (e.g., Lacey et al. 1993; Kauffmann, White, \\& Guiderdoni 1993; Cole et al. 1994; Kauffmann 1995, 1996, Baugh, Cole, \\& Frenk 1996). In the other case, the formation and evolution of galaxies is related to a gentle and coherent process of mass aggregation dictated by the forms of the density profiles of the primordial fluctuations: galaxies continuously grow inside-out. We shall call this picture, firstly developed by Gunn (1981, 1987), and by Ryden \\& Gunn (1987), the e{\\it xtended collapse scenario. }% Since disk galaxies ($\\sim 80\\%$ of present-day normal galaxies) could not have suffered major mergers due to the dynamical fragility of the disks (T\\'{o}th \\& Ostriker 1992), the extended collapse scenario results more appropriate to study their evolution. According to the merger scenario, the bulges of spiral galaxies and the elliptical galaxies arise from the mergers of galactic disks. A natural prediction of this scenario is that spirals with small bulge-to-disk ratios should have bulges older than those of spirals with large bulge-to-disk ratios (e.g., Kauffmann 1996). As Wyse, Gilmore, \\& Franx (1997) have pointed out this does not appear compatible with recent observational data (de Jong 1996a; Peletier \\& Balcells 1996; Courteau, de Jong, \\& Broeils 1997). On the other hand, if elliptical galaxies are the product of relatively recent mergers, then a big dispersion is expected in their color-magnitude relationship (but see Kauffmann 1996). Bower, Lucey, \\& Ellis (1992) showed that for the ellipticals in the Coma Cluster, this relationship is extremely tight. Ellis et al. (1997) confirmed this result for ellipticals in intermediate redshift clusters, up to $z\\sim 0.6.$ The merger scenario could also have serious difficulties from the dynamical point of view: it is not conclusive if mergers of disks are able to reproduce the high central phase-space densities of elliptical galaxies (e.g. Hernquist 1993). The inductive approach yields the possibility to establish several constraints to the galaxy formation and evolution processes. Galactic evolutionary models{\\it } have shown that due to the rapid disk gas consumption in SF, closed models are not able to explain several properties of disk galaxies, as well as the wide range of colors, gas fractions, etc. that galaxies present across the Hubble sequence (e.g., Larson \\& Tinsley 1978; Tinsley 1980; Larson, Tinsley, \\& Caldwell 1980; Kennicutt 1983; Gallagher, Hunter, \\& Tutukov 1984; Firmani \\& Tutukov 1992, 1994). On the other hand it was shown that the SF time scale in disk galaxies is not controlled by the initial gas surface density (Kennicutt 1983; Kennicutt, Tamblyn, \\& Congdon 1994). Hence, models where gas accretion is introduced are more realistic. Gas accretion could also be necessary to maintain spiral structure. In the case of open models, galaxy formation and galactic evolution might be two related processes where the SF time scale is driven by the gas accretion rate at which the disk is being built up. Infall models of disk galaxy formation have been recently favored by studies of our own Galaxy and nearby galaxies (see for references Cay\\'{o}n, Silk, \\& Charlot 1996). Moreover these inside-out disk formation models seem also to be in agreement with constrictions provided by deep field observations (e.g., Bouwens, Cay\\'{o}n, \\& Silk 1997; Cayon et al. 1996; see also Section 4). The gas infall rate in luminous galaxies may be controlled by the global process of galaxy formation (cosmological accretion) and/or by a self-regulated process of SF formation. This latter process proposed by White \\& Rees (1978) and White \\& Frenk (1991) is commonly applied in the merger scenario models. According to this mechanism, the gas accretion rate is driven by the cooling of the hot gas corona sustained by the supernova-injected energy. In the extreme situation of instantaneous galaxy formation, supernova gas reheating, halo self-regulated SF, and cooling flows (if the reheated gas was not completely expelled out of the system) become the dominant processes in regulating luminous galaxy evolution. However, the self-regulated halo SF model suffers from some inconsistencies. As Nulsen \\& Fabian (1996) pointed out, supernova feedback over large scales occurs on roughly the same time scales as the SF, not fast enough to tightly regulate the SF rate. In the same way, if a disk forms, then the self-regulating mechanism of SF will apply to the disk where other dynamical conditions prevail (see Firmani, Hern\\'{a}ndez, \\& Gallagher 1996), and not to the halo system. Unless the disk-halo connection is very effective, the SF in the disk will not be regulated by a balance of energy between the supernova input and the halo gas cooling. On the other hand, the X-ray gas corona predicted by the self-regulated SF mechanism lacks observational support, at least for the most massive galaxies for which the X-ray emission would have been above the minimum detection limits of the Rosat and ASCA experiments. The galactic infall models suggested by the inductive approach, are consistent with the cosmological (deductive) extended collapse scenario of galaxy formation and evolution. In Avila-Reese, Firmani, \\& Hern\\'{a}ndez (1997, hereafter AFH), within the framework of a standard CDM model, the MAHs corresponding to fluctuations of galactic scales were generated from the statistical properties of a Gaussian random field. After calculating the virialization of the fluctuations, a range of realistic dark halo structures were obtained. Now, with the aim to explore whether these cosmological initial conditions are able to predict the evolutionary and observational disk galaxy properties and their correlations, particularly those which go across the Hubble sequence (HS), we shall construct a self-consistent and unified model of disk galaxy formation and evolution in the cosmological context. Within the framework of the extended collapse scenario and using the galactic evolutionary models of Firmani et al. (1996), we shall study the formation and evolution of disks in centrifugal equilibrium into the evolving dark halos. In section 2, the methods we use are described. The model results at $z=0,$ the main predictions of the models, and the comparisons with observations as regards the local (\\S 3.1) and global galactic properties and their main correlations (\\S 3.2) are presented in section 3. In section 4 we compare our evolutionary models with observations at intermediate redshifts ($z\\lesssim 1)$. Finally, the concluding remarks are given in section 5. ", "conclusions": "We have modeled the formation and evolution of disk galaxies within the framework of the extended collapse scenario, which is based on the inflationary CDM models. The gas disks in centrifugal equilibrium were built-up under the assumption of detailed angular momentum conservation into spherical virializing dark matter halos whose MAHs were calculated from the initial cosmological conditions. The disk SF is produced by global gravitational instabilities and is self-regulated by an energetic balance of the turbulent gas. The bulges are formed by secular evolution of the stellar disk based on gravitational instabilities. The main predictions of the models are: 1). The disks present exponential surface brightness profiles and negative radial B-V gradients. The scale lengths and central surface brightnesses are in agreement with the observations, including the LSB galaxies. 2). The rotation curves are nearly flat up to the Holmberg radius. Contrary to observational estimations, the rotation curve decompositions show dominion of dark matter down to the galaxy central regions. A constant density core in the dark halo solves this problem. 3). The intensive properties and their correlations (particularly those which go across the HS) of the models corresponding to the local ($z\\approx 0)$ population of disk galaxies, including the LSB galaxies, are determined by the combination of three fundamental physical factors and their statistical distributions, related to the initial cosmological conditions. These three factors are the mass, the MAH, and the primordial angular momentum expressed through the spin parameter $\\lambda .$ 4). The intensive properties of the models can be described in a biparametrical sequence, where the parameters may be the color index B-V and the central surface brightness $\\mu_{B_o}$. Each one of these parameters is determined mainly by the MAH and $\\lambda $, respectively. The third fundamental physical factor, the mass, exerts no practical influence the intensive properties. We have shown that the empirical luminosity (mass)-color relation (or equivalently the color TF relation) can be explained by the effects of the metallicity and the extinction.observed dependence of extinction on luminosity (mass). These effects also contribute to decrease the slope of the B-band TF relation. 5). The SF rates of models with the average MAHs and $\\lambda =0.05$ grow by factors of 2.5-4.0 up to $z\\sim 1.5-2.5$ with respect to the SF rates at $z=0 $. After this maximum, the SF rates slowly decrease with $z.$ The SFHs of systems with early, active MAH and/or low $\\lambda ^{\\prime }s$ show high SF rates at high redfshifts ($z>3),$ while the systems with extended MAHs and/or high $\\lambda ^{\\prime }s$ present small SF rates which slowly increase until the present epoch. 6). The structural properties of the models do not change abruptly. Between $z=0$ and $z\\approx 1$ the disk scale radii in average decrease a factor $\\sim 1.3$ and the central surface brightnesses increase $\\sim 1$ mag/arcsec$^2.$ The bulge-to-total luminosity ratio also decreases with $z$ and decreases more severely for the low mass systems. The slopes of the ``structural'' and B-band TF relations do not change with $z.$ In the case of the ``structural'' TF relation, the zero-point decreases (0.75 mag for $z=0.7$ with respect to $z=0$), while for the B-band TF relation the zero-point slightly increases (0.1 mag at $z=0.7$ with respect to $z=0).$ The exploratory models presented in this work show that the main observational characteristics and correlations of disk galaxies can be well understood in the context of the extended collapse scenario, suggesting a direct connection between the conditions prevailing in the early universe and the properties of galaxies today. A serious shortcoming of galaxies emerging from Gaussian CDM cosmological models is the gravitational dominion of DM over baryon matter. The remedy to this problem is the introduction of a core in the DM halo. Fortunately, the intensive galaxy properties and their correlations are not significantly sensitive to the existence or non-existence of such a core, in such a way that all the results presented here are also true for galaxies with a core in their DM halos. The main limitations of our approach are connected to the facts that {\\it (i)} the influence of the environment on galaxy formation and evolution was not taken into account, {\\it (ii)} detailed angular momentum conservation for the baryon gas collapse was assumed, and {\\it (iii)} the mass aggregation was treated only as gas accretion neglecting the possibility of mergers of stellar systems. In future we shall address ways of overcoming these limitations with the aim to improve the model predictions and to explain galactic properties and distributions related to environment." }, "9803/astro-ph9803329_arXiv.txt": { "abstract": "Magnetic field-aligned electric fields are characteristic features of magnetic reconnection processes operating in externally agitated magnetized plasmas. An especially interesting environment for such a process are the coronae of accretion disks in active galactic nuclei (AGN). There, Keplerian shear flows perturb the quite strong disk magnetic field leading to intense current sheets. It was previously shown that given field strengths of 200 G in a shear flow, reconnection driven magnetic field aligned electric fields can accelerate electrons up to Lorentz factors of about 2000 in those objects thus providing us with a possible solution of the injection (pre-acceleration) problem. However, whereas in the framework of magnetohydrodynamics the formation of the field-aligned electric fields can be described consistently, the question has to be addressed whether the charged particles can really be accelerated up to the maximum energy supplied by the field-aligned electric potentials, since the accelerated particles undergo energy losses either by synchrotron or inverse Compton mechanisms. We pre\\-sent relativistic particle simulations starting from electric and magnetic fields obtained from magnetohydrodynamic simulations of magnetic reconnection in an idealized AGN configuration including nonthermal radiative losses. The numerical results prove that the relativistic electrons can be effectively accelerated even in the presence of an intense radiation bath. Energies from~$50 \\Dim{MeV}$ up to~$40 \\Dim{GeV}$ can be reached easily, depending on the energy density of the photon bath. The strong acceleration of the electrons mainly along the magnetic field lines leads to a very anisotropic velocity distribution in phase space. Not even an extremely high photon energy density is able to completely smooth the anisotropic pitch angle distribution which is characteristic for quasi monoenergetic particle beams. ", "introduction": "\\label{sec:intro} Active galactic nuclei (AGN) can be regarded as accreting supermassive black holes surrounded by accretion disks (Camenzind 1990; Miyoshi et al. 1995; Burke and Graham-Smith 1997 and references therein). Relativistic electrons in AGN reveal themselves by hard X-ray (probably due to pair production (cf. Svensson 1987; Done and Fabian 1989)) and $\\gamma$-ray emissions as well as radio observations of superluminous motions (e.g. Abraham et al. 1994). The $\\gamma$-radiation observed from quasars and bla\\-zars may originate in a distance $R$ of $10^{2-3}$ gravitational radii from the central engine (Dermer and Schlickeiser 1993). At that distance no significant pair production happens but relativistic leptons scatter via the inverse Compton process the IR-UV radiation of the accretion disk within a relativistically moving jet. It is well known that ``standard'' mechanisms for the acceleration of high energy leptons, as diffusive shock wave acceleration and resonant acceleration by magnetohydrodynamical (MHD) turbulence can only work efficiently for Lorentz factors larger than $\\gamma_{\\rm crit}\\simeq m_{\\rm p}/{m_{\\rm e}}$ (where $m_{\\rm p}$ and $m_{\\rm e}$ denote the proton and electron masses). Consequently, charged particles accelerated via shocks or MHD~turbulence have to be pre-accelerated which confronts us with the {\\bf injection problem} (Blandford 1994; Melrose 1994) in the AGN context. In a differentially rotating magnetized accretion disk gas, the evolution of a magnetized corona is quite hard to suppress (Galeev et al. 1979; Stella \\& Rosner 1984). Driven by the buoyancy force magnetic flux tubes ascend into the disk corona, thereby their footpoints are sheared by the differential rotation of the disk. Either by internal shear or by encountering already pre\\-sent magnetic flux, magnetic reconnection and accompanied rapid dissipation of magnetic energy happens in the coronal plasma. Such a behavior can be studied with great detail for example in the solar corona (Parker 1994 and references therein). In a recent contribution we investigated the possible role of magnetic field-aligned electric fields ($E_\\parallel $) in the context of magnetic reconnection operating in AGN coronae for the pre-ac\\-cel\\-er\\-ation of leptons (Lesch and Birk 1997; hereafter LB). It could be shown that field-aligned electric potential structures in relatively thin current sheets form. For reasonable physical parameters such electric potentials are strong enough to accelerate electrons up to $\\gamma \\approx 2000$, in principle. However, in the framework of MHD the actual energies of the accelerated particles cannot be calculated. It is the aim of the present contribution to corroborate the model introduced in LB with the help of relativistic particle simulations by taking macroscopic electric and magnetic field configurations obtained by the MHD simulations as an input. In the next section we resume the MHD model in a nutshell and present the details of the resulting three-dimensional electric and magnetic fields. In Sec.~\\ref{sec:simulation} we discuss our approach to the numerical study of high-energy particles and show the numerical results dwelling on particle spectra and energies. Eventually, we discuss our findings in Sec.~\\ref{sec:Disc}. ", "conclusions": "\\label{sec:Disc} We addressed the question of charged particle acceleration during magnetic reconnection processes by means of relativistic test particle simulations. Whereas this point is crucial for a great variety of cosmic plasmas in this contribution we dwelled on the pre-acceleration problem in the AGN context. Possible further applications include among others as different plasma systems as the terrestrial discrete auroral arcs, the solar coronal flares, radio activity in T-Tauri magnetospheres, non-thermal emission at the edges of high-velocity clouds that hit the galactic plane and the generation of electron beams in the magnetospheres of neutron stars. The starting point for our present investigations are results obtained by an MHD simulation study (LB). In contrast to previous work carried out by different other groups, we were able to study particle acceleration in large-scale non-linearly developed reconnection electromagnetic fields rather than being restricted to the prescription of somewhat idealized analytical field solutions. Moreover, since for the considered parameter regime we have to expect an intense radiation field the test particle simulations were performed including the relevant radiative losses. A very important question is whether charged particles can really be accelerated in the reconnection region up to the pretty high energy values one might deduce from the fluid treatment. Our findings indicate that, in fact, as expected from the fluid simulations (LB), leptons can be accelerated in reconnection zones located in AGN coronae up to high Lorentz factors; i.e. a significant portion of test particles gain about the maximum energy provided by the generalized field-aligned electric potential structures formed during the magnetic reconnection processes we have modeled within the framework of MHD. Thus, particle acceleration in reconnection zones may be considered as a way out of the injection problem we face in the AGN context. Since sheared magnetic fields can be expected as a very common phenomena in cosmic environments like accretion disks, stellar coronae and interstellar medium, we think that our relativistic particle studies can be regarded as realistic with respect to the used configuration and the dominant forces. In this contribution the magnetic reconnection process is driven by some resistive mechanism originating from plasma instabilities. The excited electromagnetic oscillations serve as resistance. We note that in plasmas which are collisionless (both no Coulomb collisions and no turbulent wave excitation), the particle inertia presents the ultimate source of resistivity and for the magnetic dissipation. The sheared magnetic fields in collisionless systems evolve into very thin filaments, in which the lifetime of the particle determines the electrical conductivity, thereby allowing for efficient dissipation via effective particle acceleration (Lesch and Birk 1998). Our simulations are test--particle simulations, thus, we plan for future studies to include, additionally, ponderomotive forces and the back reaction of the current carried by the high energy particles. Whether or not the latter aspect becomes important depends on the density of the run-away electrons limited by the Dreicer electric field (e.g. Benz 1993)." }, "9803/cond-mat9803258_arXiv.txt": { "abstract": "The adsorption of large ions from solution to a charged surface is investigated theoretically. A generalized Poisson--Boltzmann equation, which takes into account the finite size of the ions is presented. We obtain analytical expressions for the electrostatic potential and ion concentrations at the surface, leading to a modified Grahame equation. At high surface charge densities the ionic concentration saturates to its maximum value. Our results are in agreement with recent experiments. ", "introduction": " ", "conclusions": "" }, "9803/astro-ph9803165_arXiv.txt": { "abstract": "A method offering an order of magnitude sensitivity gain is described for using quasar spectra to investigate possible time or space variation in the fine structure constant $\\alpha$. Applying the technique to a sample of 30 absorption systems, spanning redshifts $0.51$, where $\\Delta \\alpha /\\alpha = -1.9\\pm 0.5\\times 10^{-5}$. For $z<1$, $\\Delta \\alpha /\\alpha = -0.2\\pm 0.4\\times 10^{-5}$, consistent with other known constraints. Whilst these results are consistent with a time-varying $\\alpha$, further work is required to explore possible systematic errors in the data, although careful searches have so far not revealed any. ", "introduction": " ", "conclusions": "" }, "9803/astro-ph9803023_arXiv.txt": { "abstract": "We undertake a quantitative investigation, using Monte Carlo simulations, of the amount by which quasars are expected to exceed radio galaxies in optical luminosity in the context of the `receding torus' model. We compare these simulations with the known behaviour of the \\OIII~$\\lambda$5007 and \\OII~$\\lambda$3727 emission lines and conclude that \\OIII\\ is the better indicator of the strength of the underlying non-stellar continuum. ", "introduction": "It is widely believed that radio-loud quasars and radio galaxies differ from each other only in terms of the angle between their radio axes and the line of sight (Scheuer 1987; Barthel 1989). Quasars are observed fairly close to the line of sight ($\\simlt 45\\degree$) and therefore frequently exhibit the effects of beaming, such as superluminal motion and luminous flat-spectrum radio cores, whereas radio galaxies are observed with their axes close to the plane of the sky, and do not show these effects. However, quasars also possess a luminous non-stellar optical continuum and broad emission lines, which are absent in radio galaxies. It has therefore been proposed that there is material around the nucleus which lies preferentially in the plane perpendicular to the radio axis and obscures the central regions from view in radio galaxies. Because of the geometry of this material, it is often referred to as the ``torus'', but other geometries (e.g.\\ a warped disk) are possible. Although the broad lines are hidden from direct view by this material, the narrow line region (NLR) is much larger in size and should be less strongly affected. It is therefore possible that narrow line luminosity could be an isotropic measure of the strength of the non-stellar continuum that is otherwise invisible in radio galaxies. Jackson \\& Browne (1990) tested this idea by comparing the \\OIII~$\\lambda$5007 luminosities of quasars and radio galaxies with similar redshifts and extended radio luminosities. They reported that quasars are 5--10 times more luminous in this line than radio galaxies, and attributed the difference to higher extinction in radio galaxies. Hes, Barthel \\& Fosbury (1993) performed a similar test using the \\OII~$\\lambda$3727 doublet and found that there was no measurable difference between the line luminosities of the two classes. Since \\OII\\ is at a shorter wavelength than \\OIII\\ and would be more greatly affected by a foreground screen of reddening, they suggested that most of the \\OIII\\ emission is produced close to the nucleus and is still obscured by the torus in radio galaxies, whereas \\OII\\ is produced farther out and is unaffected. Unfortunately, Jackson \\& Browne's analysis was biased because, although the radio galaxies were selected from the 3CR-based complete sample of Laing, Riley \\& Longair (1983; also called 3CRR), the quasars were drawn from a number of incomplete surveys and were therefore subject to uncertain selection effects, most notably the tendency to preferentially include optically bright objects. Since line and continuum luminosities are very well correlated in quasars, their result would be biased towards finding systematically higher line luminosities in the quasars. In addition, Jackson \\& Browne's radio galaxy sample included some objects with very weak emission lines (Class~B optical spectra; Hine \\& Longair 1979) that it is now believed may not belong to the unified scheme. An unbiased analysis significantly reduces the magnitude of Jackson \\& Browne's result, but the quasars are still about twice as luminous in \\OIII\\ than the radio galaxies. This result is also seen in the higher \\OIII/\\OII\\ ratios in broad-line objects than in narrow-line objects (e.g.\\ Saunders et al.\\ 1989; Tadhunter et al.\\ 1993). At higher redshift, however, the \\OIII\\ luminosities of the two classes of object are comparable (Jackson \\& Rawlings 1997). Rawlings \\& Saunders (1991) note that the tendency for quasars to have a higher emission line luminosity than radio galaxies could be the result of a classification bias. This could explain the discrepancy in \\OIII\\ luminosities between the two classes, but appears to run counter to the similarity in the \\OII\\ line emission. In this {\\it Letter\\/}, we apply the simple receding torus model (Lawrence 1991; Hill, Goodrich \\& Depoy 1996) to provide a quantitative explanation of the difference in \\OIII\\ luminosities between quasars and radio galaxies (and its redshift dependence), and explain why a similar effect is not seen in \\OII. The free parameters involved in our explanation are constrained by observed quantities independent of the emission line luminosities. We conclude that the luminosity of the \\OIII~$\\lambda$5007 line is an excellent indicator of the strength of the underlying non-stellar continuum. ", "conclusions": "We have used a simple receding torus model, whose free parameters are constrained by observation, to show that low redshift 3CRR quasars should be, on average, about twice as luminous in their ionizing continua as radio galaxies of the same radio luminosity. This difference should also be seen in their \\OIII, but not their \\OII, emission line luminosities, in agreement with observation. For samples with a higher quasar fraction, such as the high redshift 3CRR objects, the difference in ionizing luminosities between quasars and radio galaxies should be smaller, and there should therefore be less of a difference in their \\OIII\\ luminosities, again in line with observation. This model leads to the conclusion that the \\OIII~$\\lambda$5007 emission line, and not the \\OII~$\\lambda$3727 doublet, is an unbiased indicator of the intrinsic optical--ultraviolet luminosity of both quasars and radio galaxies." }, "9803/astro-ph9803215_arXiv.txt": { "abstract": "We report an upper limit of $9\\times 10^{12}$ cm$^{-2}$ on the column density of water in the translucent cloud along the line-of-sight toward HD 154368. This result is based upon a search for the C-X band of water near 1240 \\AA\\ carried out using the Goddard High Resolution Spectrograph of the Hubble Space Telescope. Our observational limit on the water abundance together with detailed chemical models of translucent clouds and previous measurements of OH along the line-of-sight constrain the branching ratio in the dissociative recombination of H$_3$O$^+$ to form water. We find at the $3\\sigma$ level that no more than 30\\% of dissociative recombinations of H$_3$O$^+$ can lead to H$_2$O. The observed spectrum also yielded high-resolution observations of the Mg II doublet at 1239.9 \\AA\\ and 1240.4 \\AA, allowing the velocity structure of the dominant ionization state of magnesium to be studied along the line-of-sight. The Mg II spectrum is consistent with GHRS observations at lower spectral resolution that were obtained previously but allow an additional velocity component to be identified. ", "introduction": "Translucent clouds have total extinctions $A_{\\rm V}$ in the range of 2-5 magnitudes. The physical and chemical conditions that characterize translucent clouds are therefore intermediate between those in diffuse and in dense molecular clouds (Crutcher 1985; van Dishoeck \\& Black 1988). Photodissociation rates are significantly smaller in translucent clouds than in diffuse clouds, and the column densities of molecules like CO, OH and CS are correspondingly larger. Translucent clouds can be observed through absorption lines of CN, CH, CH$^+$ and C$_2$ as well as through millimeter emission lines of CO and CS. Although they are an abundant constituent of dense molecular clouds (Jacq et al.\\ 1988; Zmuidzinas et al.\\ 1995; Gensheimer et al.\\ 1996; van~Dishoeck \\& Helmich 1996; Cernicharo et al.\\ 1997) and are expected to be the dominant coolant of warm dense regions within such clouds (Neufeld \\& Kaufman 1993; Neufeld, Lepp \\& Melnick 1995), water molecules have not been detected in diffuse or translucent molecular clouds. In this paper, we report the results of a search for water in the translucent cloud along the line of sight to HD 154368. HD 154368 is an O9.5 Iab star at an estimated distance of 800 pc from the Sun (Snow et al.~1996) that is situated near the Sco OB 1 association (Blades \\& Bennewith 1973). The line-of-sight toward HD 154368 lies about $\\rm 14^o$ from the core of the $\\rho$ Oph molecular cloud, which is located at a distance of 125$\\pm$25 pc (de Geus, de Zeeuw, \\& Lub 1989) and has a heliocentric radial velocity of $\\rm -6.6\\, km \\,s^{-1}$. The main H I (e.g.\\ Riegel \\& Crutcher 1972) and Na I (e.g.\\ Crawford, Barlow, \\& Blades 1989) absorption features are observed at a radial velocity that is close to that of the $\\rho$ Oph molecular cloud, a result that suggests that the main component of gas toward HD 154368 is probably not close to the star and is more likely to be the outer envelope of a dense molecular cloud only about 125 pc from the Sun. The color excess along the line of sight is E(B--V)=0.82 and most of the gas toward HD 154368 resides in two clouds centered near --3.26 (main component) and --20.95 km s$^{-1}$ (heliocentric). This line-of-sight has been observed extensively from the ground by means of narrow H I 21 cm absorption features (e.g.~Riegel \\& Crutcher 1972); optical absorption lines of Na~I~D (e.g.~Crawford et al.~1989), CN B--X (0,0) and A--X (0,0), CH, CH$^+$; and near-infrared absorption lines of C$_2$ in the A--X Phillips system at 8750 \\AA\\ (van Dishoeck \\& de Zeeuw 1984). The CH observations can be used to infer the total column density of H$_2$ along the line-of-sight (Gredel, van Dishoeck, \\& Black 1993), a quantity that is needed to determine the atomic depletions. The red system of CN has been used by Gredel, van Dishoeck, \\& Black (1991) to derive an electron density of 0.05-0.15 cm$^{-3}$ for the molecular component. Molecular emission lines are also detectable toward HD 154368. Data on $^{12}$CO $J=1-0$, $J=2-1$, and $J=3-2$ and $^{13}$CO $J=1-0$ have been presented by van Dishoeck et al.~(1991), and have been used to constrain the column density of CO and the density of H$_2$ in the molecular component. These authors confirmed the relatively low density $n_{\\rm H}\\approx 350$ cm$^{-3}$ derived independently from the C$_2$ absorption data. The $^{12}$CO $J=1-0$ distribution over a region of $30'\\times 30'$ around the star is featureless, a result that suggests once more that the cloud and the star are not located close to one another. The velocity structure of the gas toward HD 154368 is well known through the Na I data obtained using the Ultra High Resolution Facility at the Anglo Australian Telescope (Snow et al.~1996) and high resolution Ca II observations (Crawford 1992). The Na I data indicates seven velocity components at --27.7, --20.95, --18.2, --14.75, --10.5, --3.26, and +5.62 km s$^{-1}$ where 96\\% of the neutral sodium resides in the --3.26 km s$^{-1}$ feature. The Ca II results show five velocity components at --27.6, --20.9, --14.4, --4.3, and +6.8 km s$^{-1}$. These five coincide roughly with the Na I components, but their relative column densities are different. Approximately 50\\% of the ionized calcium resides in the --4.3 km s$^{-1}$ feature with an even distribution over the remaining velocity components. A detailed discussion of all these observations and the implications they have for the physical conditions of the ambient medium can be found in Snow et al.~(1996). In this work, their results will be adopted and the main focus will be on the resulting OH and H$_2$O chemistry. ", "conclusions": "We have reported a $3\\sigma$ upper limit of $9\\times 10^{12}$ cm$^{-2}$ on the column density of water toward HD 154368. We have constructed detailed chemical models which incorporate many existing constraints on the physical conditions of the ambient medium. Together with the known column density of OH along the line of sight, we constrain the H$_2$O branching ratio for dissociative recombination of H$_3$O$^+$ to be smaller than 30\\%. Our results are in agreement with the laboratory studies of Williams et al.\\, which find an oxygen channel in the recombination of H$_3$O$^+$ and a small branching ratio for the production of water, but are mildly inconsistent with the laboratory results of Vejby-Christensen et al. The observed spectrum of HD 154368 also yielded high-resolution observations of the Mg II doublet at 1239.9 \\AA\\ and 1240.4 \\AA, allowing the velocity structure of the dominant ionization state of magnesium to be studied along the line-of-sight. The Mg II spectrum is consistent with GHRS observations at lower spectral resolution that were obtained previously but allow an additional velocity component to be identified. Absorption by the Ge II line at 1237.059 \\AA\\ was also detected, the first detection of interstellar germanium along the line-of-sight to HD 154368. We acknowledge with gratitude the support of NASA grant NAGW-3147 from the Long Term Space Astrophysics Research Program and of grant GO-06739.01-95A from the HST Cycle 6 Program. In the final stages of this project MS was partially supported by NASA through grant HF-01101.01-97A, awarded by the Space Telescope Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA under contract NAS 5-26555. Most of the simulations presented in this work were performed on the Cray YMP operated by the Netherlands Council for Supercomputer Facilities in Amsterdam. \\newpage" }, "9803/astro-ph9803287_arXiv.txt": { "abstract": "ASCA observations of NGC 4636 and a southern region have revealed extended X-ray emission to a radius of about 300 kpc from the galaxy. The symmetric nature of the observed surface brightness around NGC 4636 indicates its association to this galaxy rather than to the Virgo cluster. Model independent estimation of the gravitational mass profile shows a flattening at a radius of $20 \\sim 35$ kpc, where the total mass reaches $\\sim 6\\times 10^{11} M_\\odot$ and a mass-to-light ratio of 23. The mass still increases to larger radii, reaching $9\\times10^{12} M_\\odot$ with a mass-to-light ratio of 300 at $\\sim$ 300 kpc from NGC~4636. These features suggest presence of a galaxy group surrounding NGC 4636. If such optically dark groups are common among X-ray bright ellipticals, it would explain the very large scatter in their X-ray luminosities with similar optical luminosities. ", "introduction": "Spatial distribution of gravitational mass around elliptical galaxies has been estimated from radial distribution of X-ray emitting hot interstellar medium (ISM) (e.g. \\cite{3}; \\cite{4}). However, the outermost radius of the ISM distribution has so far remained undetermined due to a lack of sensitivity. As a result, the total gravitating masses of elliptical galaxies depend heavily on the assumed overall extent of the ISM\\@. High sensitivity X-ray observations of bright galaxies are therefore important in determing the mass distribution in these systems. NGC~4636 is one of the nearest giant elliptical galaxies. We employ a distance of 17 Mpc (\\cite{8}). Diffuse thermal X-rays from its ISM make it one of the X-ray brightest elliptical galaxies (e.g. \\cite{3}; \\cite{4}) except those located at the center of clusters. Although NGC~4636 is located in the Virgo Southern Extension (\\cite{nol}), it is apparently clear of the large-scale X-ray emission from hot intra-cluster gas associated with the Virgo cluster (\\cite{10}; \\cite{11}). Trinchieri et al.\\ (1994), using {\\it ROSAT}, detected a largely extended X-ray emission up to 18$'$ from NGC~4636, and suggested it to be an extension of the Virgo cluster emission. Here, we present new results on the extended X-ray emission around NGC~4636 based on a very long observation towards the galaxy center and on an offset observation from ASCA. Spectral properties of NGC~4636 have been already reported by Matsushita et al.\\ (1997). ", "conclusions": "ASCA observations have shown a very extended X-ray emission with a radius of at least $60'$ (300 kpc) surrounding NGC~4636, which is much larger than that detected by {\\it ROSAT}\\@. We studied radial temperature and density distributions of the gas, and determined gravitational mass profile with and without assuming a double-$\\beta$ model. According to Fig.\\ 3, the gravitational halo of NGC~4636 appears to terminate once at $R \\sim 10-30$ kpc, where the total mass reaches several $ \\times 10^{11} ~ M_\\odot$. This inner component probably corresponds to the mass associated with the galaxy NGC~4636. Then, the total mass starts increasing again, forming a halo-in-halo structure whose total size is comparable to that of a galaxy group. The mass of the whole system and its mass-to-light ratio at 300 kpc reach $\\sim 1 \\times 10^{13}~ M_\\odot$ and $\\sim 300$, respectively, which are again comparable to those of galaxy groups (\\cite{28}). At $R \\ge 200$ kpc, the X-ray emitting plasma becomes the dominant form of baryons, with its mass reaching 5--8\\% of the total gravitating mass. This value is considerably higher than those of individual elliptical galaxies, and again close to those found in galaxy groups and poor clusters (\\cite{28}). These features all support the presence of a gravitational potential with a size of a galaxy group around NGC~4636. A further support is given by the observed abundance decrease in the X-ray emitting plasma (\\cite{12}; \\cite{15}; \\cite{matusita97}; \\cite{17}), from $\\sim 1$ solar within $\\sim 30$ kpc, to $\\sim 0.2$ solar beyond $\\sim 50$ kpc to a is typical level of groups and clusters of galaxies. Nolthenius (1993) identified % NGC~4636 as a member of Virgo Southern Extension F Cloud. However the 3-dimensional location of NGC~4636 is far offset from the center of the member galaxies in the cloud, and their velocity dispersion, 463 km s$^{-1}$, is much larger than that inferred from the observed gas temperature. The diameter of the cloud, 1.26 Mpc, is also much larger than the scale of the X-ray emission. Therefore, the relation between the X-ray halo and the galaxy cloud remains unclear, but it is very likely that there is some galaxy concentration around NGC~4636. Extensive studies of ASCA data have revealed (\\cite{17}) that X-ray luminous elliptical galaxies preferentially possess large-scale X-ray halos of a few 100 kpc scale, just like NGC~4636. Such a galaxy may be regarded as a dominant member of a galaxy group even though its evidence is scarce in the optical data. Thus, the X-ray emitting plasma may provide a better tracer of the total mass distribution than the light emitting matter. We speculate that X-ray luminous elliptical galaxies may commonly possess such an extended emission with low surface brightness. X-ray luminosity of elliptical galaxies scatter by 2 orders of magnitude for the same optical luminosity, and its cause has long been a puzzle. The presence and absence of the extended emission can easily account for the large scatter in the total X-ray luminosity, and the low surface brightness of the extended emission, such as in NGC~4636, explains the previous undetection. We hope that systematic deep exposures from ASCA will solve this problem and bring us a complete understanding of the mass distribution around elliptical galaxies. \\medskip K.M. acknowledges support by the Postdoctoral Fellowship of the Japan Society for Promotion of Science. \\clearpage" }, "9803/astro-ph9803078_arXiv.txt": { "abstract": "Recent hard X-ray spectroscopy of active galactic nuclei has strongly suggested that double-peaked, very broad Fe K emission arises from an accretion disk around the central engine. Model fitting of the observed Fe K emission line profile makes it possible to estimate a probable inclination angle of the accretion disk. In order to study the geometrical relationship between the accretion disk and broad emission-line regions (BLRs), we investigate the correlation between the inclination angle of the accretion disk and the velocity width of BLRs for 18 type-1 Seyfert galaxies. We found that there may be a negative correlation between them, i.e., Seyfert nuclei with a more face-on accretion disk tend to have larger BLR velocity widths, suggesting that the BLRs are not coplanar with respect to the accretion disk. The most probable interpretation may be that the BLRs arise from outer parts ({\\it r} $\\sim$ 0.01 pc) of a warped accretion disk illuminated by the central engine. ", "introduction": "Given the current paradigm of active galactic nuclei (AGNs), the observed huge luminosities of AGNs are powered by gravitational accretion onto a supermassive black hole (e.g., Rees 1984; Blandford 1990; Antonucci 1993; Peterson 1997). The central engines are considered to be surrounded by dusty tori whose typical inner radii are on the order of $\\sim$ 1 pc (e.g., Antonucci, Miller 1985; Pier, Krolik 1992, 1993). Therefore, in order to understand AGNs, it is very important to investigate the spatial structures of the inner $\\sim$ 1 pc regions in which a supermassive black hole, an accretion disk, warm absorbers, and broad emission-line regions (BLRs) reside. The innermost constituent is the accretion disk; its typical radius is $\\sim$ (10 --- 100) $R_{\\rm S} \\sim (10^{-4}$ --- $10^{-3})M_{8}$ pc where $R_{\\rm S}$ is the Schwarzschild radius of a black hole and $M_{8}$ is the black-hole mass in units of $10^{8} M_{\\sun}$. The existence of accretion disks in AGNs has been demonstrated by the recent X-ray spectroscopy, collected using {\\it ASCA} (Tanaka et al. 1994). The {\\it ASCA} X-ray spectra of type-1 AGNs show the presence of very broad Fe K$\\alpha$ emission, whose line profile can be fitted well by some accretion-disk models (Tanaka et al. 1995; Fabian et al. 1995; Mushotzky et al. 1995; Iwasawa et al. 1996; Nandra et al. 1997; Reynolds 1997). An ionized accretion disk has also been detected by the recent radio continuum mapping in the archetypical, nearby Seyfert galaxy NGC 1068 (Gallimore et al. 1997). Another inner constituent is broad-line regions (BLRs), because their typical radii are of on the order of 0.01 pc (e.g., Peterson 1993). One of the most important questions related to the BLRs is how emission-line clouds are distributed around AGNs. Although there is still no definite consensus concerning the dynamical and spatial structure of BLRs, there are three alternative models: 1) the disk model (Shields 1979; see also Osterbrock 1989), 2) the high-velocity streamer model (Zheng et al. 1991), and 3) a pair of conical regions in which photoionized clouds are orbiting randomly with Keplerian motion (Wanders et al. 1995; Wanders, Peterson 1996, 1997; Goad, Wanders 1996). In particular, a recent detailed analysis of the reverberation mapping has strongly suggested that the BLRs of the type-1 Seyfert galaxy NGC 5548 has the third type of geometry (Wanders et al. 1995). It is, however, still not known which model is the most popular one. Since recent observations have shown that the BLRs are dominated by rotational motion, rather than the radial motion (e.g., Peterson 1993; Wanders et al. 1995), it is interesting to examine whether or not the rotational axis of the BLR is nearly the same as that of the accretion disk. If the disk model for BLRs would be correct, the BLRs may be coplanar with respect to the accretion disk. In fact, double-peaked BLRs have sometimes been considered to arise from accretion disks, themselves [e.g., P\\'erez et al. (1988); Livio, Xu (1997) and references therein; see also, however, Gaskell (1996)]. Motivated by this, we investigate the relationship between the inclination angle of the accretion disk and the width of BLR statistically using published data. ", "conclusions": "The most important result in this study is that there is no obvious {\\it positive} correlation between $\\sin i_{\\rm AD}$ with FWZI(H$\\alpha$)/ $L^{1/4}_{\\rm X}$. This suggests that {\\it the BLRs are not coplanar with respect to accretian disks}. Some Seyfert nuclei in our sample show double-peaked BLRs (DBLRs). It has sometimes been considered that such DBLRs may arise from an accretion disk, itself (e.g., P\\'erez et al. 1988). The DBLR emission profiles of the four Seyfert nuclei in our sample (NGC 3227, NGC 3783, NGC 5548, and 3C 120) were studied by Rokaki et al. (1992) using a standard geometrically-thin accretion- disk model; also, the inclination angles of the DBLRs ($i_{\\rm BLR}$) were derived. We compare these inclination angles with those of the accretion disks in figure 2. This comparison also suggests a negative correlation between $i_{\\rm AD}$ and $i_{\\rm BLR}$, being consistent with our result. This strengthens our suggestion that the BLRs are not coplanar with respect to the accretion disks in the Seyfert nuclei studied here. Let us consider what kind of geometrical configuration can explain the non-coplanar property. The negative correlation means that the normalized velocity width increases with decreasing inclination angle; i.e., {\\it Seyfert nuclei with a more face-on accretion disk tend to have larger BLR velocity widths}. There may be three alternative ideas to explain this property. One is the bipolar streamer model (e.g., Zheng et al. 1990). If we observe the accretion disk from a face-on view, the velocity width would be widest because the bipolar wind flows along our line of sight. However, this model has an intrinsic difficulty, as claimed by Livio and Xu (1996), because the emitting region on the receding flow (jet) is obscured from view by the accretion disk; the standard, optically thick accretion disk is opaque up to $\\sim$ 1 pc, and, thus, the BLR component behind the disk cannot be seen, because the typical radial distance of BLRs from the central engine is on the order of 0.01 pc (e.g., Peterson 1993). The second idea is that BLRs are located in nearly the same plane as that of an accretion disk, but are orbiting with poloar orbits. If a two-sided jet is ejected with a highly inclined angle with respect to the {\\it global} accretion disk, we can explain the negative correlation. Such a jet model is briefly described by Norman and Miley (1984). This idea is consistent with the recent reverberation mapping result for the BLRs of NGC 5548 because the most likely geometry of the BLRs of this galaxy is a pair of conical regions in which photoionized clouds are orbiting randomly with Keplerian motion (Wanders et al. 1995; Wanders, Peterson 1996, 1997; see also Goad, Wanders 1996). This model may also have the same obscuration problem as that for the above streamer model. However, if the BLR clouds are moving at randomly oriented orbits (Wanders et al. 1995), there may be no obscuration problem. The third idea is that BLRs arise from outer parts of a warped accretion disk. The disk model for BLR is the standard idea (Shields 1977; see also for a review Osterbrock 1989). It has been recently shown that accreting gas clouds probed by water-vapor maser emission at 22 GHz show evidence of significant warping (Miyoshi et al. 1995; Begelman, Bland-Hawthorn 1997). The warping of accretion disks can be driven by the effect of the radiation-pressure force (Pringle 1996, 1997). For typical AGN, the warping may occur at {\\it r} $>$ 0.01 pc (Pringle 1997), which is at a similar distance as BLRs. Therefore, the warped-disk model can explain the observed negative correlation reasonably well. This model is schematically shown in figure 3. Since the degree of warping and the viewing angle are different from AGN to AGN, the negative correlation between $i_{\\rm AD}$ and $i_{\\rm BLR}$ may be blurred as obtained in our analysis, although the poor correlation may be also due to the large errors in the estimate of $i_{\\rm AD}$." }, "9803/astro-ph9803308_arXiv.txt": { "abstract": "We present measurements of the oxygen abundances in 64 H{\\sc ii} regions in 12 LSB galaxies. We find that oxygen abundances are low. No regions with solar abundance have been found, and most have oxygen abundances $\\sim 0.5$ to 0.1 solar. The oxygen abundance appears to be constant as a function of radius, supporting the picture of quiescently and sporadically evolving LSB galaxies. ", "introduction": "Low surface brightness (LSB) disk galaxies have all the characteristics of unevolved galaxies. Those discovered so far constitute a population of gas-rich, metal-poor galaxies with very low star formation rates (see the review by Bothun et al. 1997). Their surface brightnesses are a few magnitudes lower than the values commonly found for so-called normal galaxies. Most of them are rather late-type galaxies, with diffuse spiral arms. A direct probe of the evolutionary state of these galaxies is the metal abundance in the interstellar medium (ISM). A low abundance generally indicates only limited enrichment of the ISM and therefore (in a closed system) a small amount of evolution. Because of their low surface brightness, obtaining spectra of the stellar disks of LSB galaxies is difficult and requires large amounts of telescope time. Conclusions on metallicities must therefore be derived from spectra of H{\\sc ii} regions. These usually are the brightest distinct objects in a LSB galaxy. Their bright emission lines make them more easily observable than the underlying continuum. The H{\\sc ii} regions that are observed in LSB galaxies are usually giant H{\\sc ii} regions, that are ionized by star clusters rather than by a few stars. The first measurements of the oxygen abundances in H{\\sc ii} regions in LSB galaxies were presented in McGaugh (1994). He found, using an empirical oxygen abundance indicator, that LSB galaxies are low metallicity galaxies with typical values for the metallicity $Z<\\frac{1}{3} Z_{\\odot}$. It shows that low metallicities can occur in galaxies that are comparable in size and mass to the bright galaxies that define the Hubble sequence. As LSB galaxies are found to be isolated (Mo et al.\\ 1994), this suggest that surface mass density and environment are as important for the evolution of a galaxy as total mass. In this paper we present a follow-up study of oxygen abundances in LSB galaxies. We confirm the results by McGaugh (1994) that LSB galaxies are metal-poor. We present two direct measurements of the oxygen abundance from measurements of the [O {\\sc iii}]$\\lambda 4363$ line, supplemented with a large number of empirically determined oxygen abundances. In addition, for those galaxies where sufficient data are available, we investigate the change in abundance with radius, and show that the measurements are consistent with no gradient. The steeper gradients found in HSB galaxies (Vila-Costas \\& Edmunds 1992 [VCE], Zaritsky et al. 1994, Henry \\& Howard 1995, Kennicutt \\& Garnett 1996) are not present. It is worth noting that the exact form and magnitude of the Milky Way oxygen gradient has now been consistently reproduced in early-type stars, H{\\sc ii} regions and planetary nebulae, which supports that the extra-galactic H{\\sc ii} region gradients in HSB galaxies are real (Smartt \\& Rolleston 1997). The lack of abundance gradients in LSB galaxies supports the picture of stochastic and sporadic evolution, where the evolutionary rate only depends on local conditions and not on the global properties of LSB galaxies as a whole. Section 2 describes the sample selection and observations. Section 3 presents the data, while in Sect.~4 the analysis is described. Section 5 discusses the abundances found. Section 6 discusses reddening towards the H{\\sc ii} regions, while Sect.~7 concludes with presenting the gradients. In Sect.~8 the results are summarized. ", "conclusions": "The oxygen abundances in 64 H{\\sc ii} regions in 12 LSB galaxies have been measured. Oxygen abundances are low. No region with solar abundance has been found, and most have oxygen abundances that are $\\sim 0.5$ to 0.1 solar. No strong radial oxygen abundance gradients are found. The abundance seems to be constant, rather, as a function of radius, supporting the picture of quiescently and sporadically evolving LSB galaxies." }, "9803/hep-ph9803471_arXiv.txt": { "abstract": "\\noindent From a phenomenological point of view, we study active-active and active-sterile flavour-changing (and flavour-conserving) oscillations of Dirac-Majorana neutrinos both in vacuum and in matter. The general expressions for the transition probabilities in vacuum are reported. We then investigate some interesting consequences following from particular simple forms of the neutrino mass matrices, and for the envisaged scenarios we discuss in detail neutrino propagation in matter. Special emphasis is given to the problem of occurrence of resonant enhancement of active-active and active-sterile neutrino oscillations in a medium. The peculiar novel features related to the Dirac-Majorana nature of neutrinos are particularly pointed out. ", "introduction": "Today we have several indications in favour of non zero neutrino masses and mixing.\\\\ The solar neutrino problem, i.e. the observed deficit of solar neutrino fluxes \\cite{SNP}, is a well established tool whose resolution requires (almost without doubt \\cite{doubt}) neutrino physics beyond the (minimal) Standard Model. The acceptable solutions to this problem, in terms of vacuum \\cite{Vsol} or matter \\cite{Msol} flavour oscillations or spin and flavour oscillations \\cite{SFsol} as well as in terms of active-sterile neutrino conversion \\cite{Stsol}, all need non vanishing neutrino masses and mixing \\cite{Vac, MSW, Akh, sterile}.\\\\ The second indication in favour of neutrino oscillations come from the observed deficit of atmospheric muon neutrinos with respect to electron neutrinos \\cite{atmo} that can be explained in terms of \\nm \\rt \\nt or \\nm \\rt \\ne or even active-sterile neutrino transitions \\cite{atsol}.\\\\ Laboratory direct searches for massive neutrinos only give, at present, upper limits on neutrino masses \\cite{masses} and the same is valid for reactor and accelerator neutrino oscillations experiments \\cite{reactor}, except for the LSND experiment \\cite{LSND} whose results seem to be explained in terms of $\\ov{\\nm} \\rt \\ov{\\ne}$ oscillations.\\\\ Hints for massive neutrinos also come from cosmology, looking at \\nt as the most probable candidate for the hot component of the dark matter \\cite{HDM} and from the observed peculiar velocities of pulsars \\cite{Segre}. On the other hand, in Grand Unified Theories, which attempt to give a unified view of electroweak and strong interactions, massive neutrinos are predicted \\cite{Buccella} together with other phenomena violating both lepton numbers and baryon number (such as, for example, proton decay). However, the most intriguing fact is that a simple scenario with only three massive neutrinos cannot account for the solar neutrino problem, the atmospheric neutrino anomaly and the LSND result. This is because the three squared masses differences $\\Delta m^2$ for the three oscillation solutions to these problems are all distinct between them: the resonant MSW solution to the solar neutrino problem requires $\\Delta m^2 \\, \\sim \\, 10^{-5}$ eV$^2$, while for the atmospheric anomaly $\\Delta m^2 \\, \\sim \\, 10^{-2}$ eV$^2$ is needed and $\\Delta m^2 \\, \\sim \\, 1$ eV$^2$ for the LSND result. Many analyses \\cite{analyses} have been conducted for giving a unified view of the three problems in terms of neutrino oscillation (taking into account also the limits from laboratory experiments) and a coherent picture seems to emerge with four massive neutrinos, namely the three known neutrinos plus a sterile neutrino. Note that four neutrino mass eigenstates are needed, but not necessarily more than three neutrino flavour eigenstates. This scenario is easily realized if one considers neutrinos as Dirac-Majorana particles described by the following general mass term in the electroweak lagrangian \\cite{sterile}: \\be -{\\cal L}^{DM}_{m} \\; = \\; \\sum_{l,l^{\\prime}} \\ov{\\nu}_{l^{\\prime}R} \\; M^{D}_{l^{\\prime}l} \\; \\nu_{l^{\\prime}L} \\; + \\; \\frac{1}{2} \\; \\sum_{l,l^{\\prime}} \\ov{\\nu}^{c}_{l^{\\prime}R} \\; M^{1}_{l^{\\prime}l} \\; \\nu_{l L} \\; + \\; \\frac{1}{2} \\; \\sum_{l,l^{\\prime}} \\ov{\\nu}^{c}_{l^{\\prime}L} \\; M^{2}_{l^{\\prime}l} \\; \\nu_{l R} \\; + \\; h.c. \\label{11} \\ee Here $l,l^{\\prime} = e, \\mu , \\tau$ label the three flavour eigenstates and $M_D$, $M_1$, $M_2$ are the Dirac and the two Majorana mass matrices which, in general, are hermitian and non diagonal (however, $M_1$ and $M_2$ have to be symmetric). To construct the mass term in (\\ref{11}) we need the three known left-handed neutrinos (and their antiparticles) and other three right-handed sterile neutrinos (and their antiparticles) \\footnote{Obviously, the generalization to more than three families is possible and straightforward}. After the diagonalization of (\\ref{11}) we can obtain in general six mass eigenstates which are Majorana fields; so in this framework we can easily endow the above scenario with four massive neutrinos coming from the experiments.\\\\ Note that if neutrinos are really described by the mass term in (\\ref{11}), the total lepton number is no longer conserved and peculiar phenomena, as neutrinoless double beta decay and neutrino-antineutrino oscillations can take place.\\\\ We stress that (\\ref{11}) is predicted in many GUTs \\cite{Buccella} in which the popular ``seesaw'' mechanism \\cite{seesaw} can give rise to very small neutrino masses in a very natural way by supposing $M_1 \\approx 0$ and $M_D \\ll M_2$. However, this is not the only possibility; recently some models assuming $M_1 \\simeq M_2$ have been proposed \\cite{equal} for accounting the three experimental indications on neutrino oscillations discussed above. Here we further explore this last scenario and study flavour transitions of Dirac-Majorana neutrinos from a completely phenomenological point of view, adopting no particular model. This work is a generalization to flavour transitions of previous papers \\cite{TE, previous} in which we studied peculiar oscillations of Dirac-Majorana neutrinos. We now assume, for simplicity, only two flavours, so $M_D$, $M_1$, $M_2$ in (\\ref{11}) are $2 \\times 2$ matrices in the flavour space. In the following section, the basic vacuum oscillations allowed by (\\ref{11}) are studied and transition probabilities are explicitly given in the general case. Some very interesting consequences due to particularly simple forms of the mass matrix are also investigated. In section 3, given the effective hamiltonian of Dirac-Majorana neutrinos interacting with a medium, resonant matter oscillations are considered along with a qualitative discussion of the phenomenon with the aid of the level crossing diagram. Finally, in section 4, there are our conclusions and remarks. ", "conclusions": "In this paper we have studied the propagation both in vacuum and in matter of Dirac-Majorana neutrinos and analyzed active-active (flavour-changing) as well as active-sterile transitions, which are, in general, both possible.\\\\ For vacuum oscillations, in section 2 we have given the general expressions for the transitions probabilities for $ \\nel \\rightarrow \\nml $, \\nel \\rt \\ncml , $ \\nel \\rightarrow \\ncel $ We have then discussed some interesting limiting cases for Dirac ($M_D$) and Majorana ($M_M$) mass matrices. For pure Dirac ($M_M=0$) or pure Majorana ($M_D = 0$) neutrinos obviously we recover the usual flavour oscillation formulae \\cite{Vac}, while for both $M_D$ and $M_M$ non vanishing and diagonal the Pontecorvo formula \\cite{Pontecorvo} for neutrino-antineutrino (active-sterile) oscillations is obtained \\cite{TE}.\\\\ An interesting non trivial case is that with $M_D$ and $M_M$ given by (\\ref{226}) or (\\ref{235}) which implement the idea that neutrino mixing is essentially ruled only by the Dirac mass term while the Majorana mass term is diagonal or vice-versa, respectively. In both cases, neither pure flavour oscillations nor Pontecorvo oscillations are predicted, but only flavour-changing active-sterile transitions, such as $\\nel \\rightarrow \\ncml$, are possible. Remarkably, the transition probability for these is identical in form to that for flavour oscillations for pure Dirac or Majorana neutrinos, and this holds both in vacuum and in matter. For the latter, the resonance condition is only shifted by the neutral current contribution of \\nel to the effective potential. So, for example, the solution to the solar neutrino problem in terms of active-sterile neutrino oscillations proposed in \\cite{Stsol} applies unmodified to the present scheme. \\\\ Another interesting case, even if a bit more complicate, has been analysed for the mass matrices in (\\ref{236}) or (\\ref{237}), which implements the idea that neutrino mixing is given by the Dirac and Majorana mass terms with the same strength. In this case, $ \\nel \\rightarrow \\nml $, $ \\nel \\rightarrow \\ncml $, $ \\nel \\rightarrow \\ncel $ transitions are all possible and, in vacuum, the first two have the same transition probability, which is also identical in form to that obtained in the previous case, except for a constant suppression factor in the amplitude of oscillations of 1/4. Also in matter the pattern of neutrino transitions present in the general case is (qualitatively) reproduced in this peculiar scheme. In particular, all the flavour changing oscillations can be resonantly amplified while Pontecorvo \\nel \\rt \\ncel matter oscillations have maximum amplitude only for a given electron to neutron number ratio (see eq. (\\ref{313}) and the related footnote); the resonance conditions were discussed in section 3.1 . \\\\ Given the multiresonance structure of the oscillations pattern, it is then interesting to follow the evolution of a \\nel , for example, in a varying density medium such as the Sun; this has been done in section 3.4 with the help of a level crossing diagram reported in Fig. 1. Several scenarios are possible according to the adiabaticity properties of level crossing near the resonance points. In particular, starting from a pure \\nel beam at high density, to have a consistent conversion into \\nml at low density the resonance for \\nel \\rt \\ncml has to be crossed non adiabatically, while the passage through the one for \\nel \\rt \\nml has to be adiabatic. However, we have also shown that at very low density, and then in the vacuum, it is more appropriate to deal with the Majorana combinations $\\wt{n_\\pm}$ in (\\ref{212}) than with the pure flavour states \\nel , \\ncel , \\nml , \\ncml . This is strictly related to the Dirac-Majorana nature of neutrinos, which chooses Majorana eigenstates instead of pure flavour states as starting points. In this respect, we have to deal with ``generic'' flavour-changing or flavour-conserving transitions of Dirac-Majorana neutrinos without looking at the particular active neutrino or sterile antineutrino state. It is through the weak interactions, with which neutrinos are produced and detected, that a particular (active or sterile) component of the Majorana eigenstates is chosen.\\\\ The results obtained for the case of degenerate Dirac-Majorana mixing are qualitatively valid also in the general case in which all the entries of the Dirac and Majorana mass matrices are non zero and different between them. The main difference between the two cases is that in the general framework there are 3 mass parameters and two mixing angles ruling the evolution, while for the particular case studied in sections 2.3 and 3.4 there are only 2 mass parameters and 1 mixing angle (these parameters being not completely independent, because of relation (\\ref{239b})). The presence of more degrees of freedom in the general case allows to consider some peculiar situations which are not possible otherwise. The most remarkable one is that in the general case the proportionality of the $\\Sigma$ parameter to $\\sin^2 \\, 2 \\theta_+$ (see eq. (\\ref{239b})) is lost, so that the structure of the level crossing diagram at very low density can be altered. The eigenvalues of $H_m$ in (\\ref{39}) for zero density (vacuum) are given by \\be \\frac{1}{8k} \\, \\left( \\pm \\, 2 \\Delta m_+^2 \\; + \\; \\Sigma \\right) \\ee \\be \\frac{1}{8k} \\, \\left( \\pm \\, 2 \\Delta m_-^2 \\; - \\; \\Sigma \\right) \\ee so that one can manipulate the mass parameters to modify the low density region of the level crossing diagram without grossly altering the region where the resonance points are present. In any case, there can be present no substantial modifications of the conclusions reached above. The oscillations of Dirac-Majorana neutrinos here studied with their peculiar features can be efficiently tested in astrophysics, in particular detecting solar or supernova neutrinos, and can have even profound implications in cosmology for the nucleosynthesis of light elements in the Universe. \\vspace{1truecm} \\noindent {\\Large \\bf Acknowledgements}\\\\ \\noindent We express our sincere thanks to Prof. F. Buccella for very useful talks and his unfailing encouragement, and to Prof. E. Kh. Akhmedov for enlightening discussions with one of us (S.E.)." }, "9803/astro-ph9803234_arXiv.txt": { "abstract": "A new cataclysmic variable is identified as the optical counterpart of the faint and hard X-ray source RX\\,J0757.0+6306 discovered during the ROSAT all-sky survey. Strong double-peaked emission lines bear evidence of an accretion disc via an S--wave which varies with a period of 81$\\pm 5$~min. We identify this period as the orbital period of the binary system. CCD photometry reveals an additional period of 8.52$\\pm 0.15$ min. which was stable over four nights. We suggest that \\rxj\\ is possibly an intermediate polar, but we cannot exclude the possibility that it is a member of the SU UMa group of dwarf novae. ", "introduction": "Within a project for the optical identification of a complete sample of 674 northern ROSAT All-Sky Survey (RASS) X-ray sources (which is a collaboration between the Max-Planck-Institut f\\\"ur extraterrestrische Physik, Garching, the Landessternwarte Heidelberg, Germany, and the Instituto Nacional de Astrof\\'isica, Optica y Electronica, Mexico) several new cataclysmic variables were identified. A detailed description of the project is given by Zickgraf \\etal\\ (1997). The full catalogue with all identifications is published in Appenzeller \\etal\\ (1998). Here we report the identification of the RASS X-ray source RX\\,J0757.0+6306 (= 1RXS J075700.5+630602). Cataclysmic variables (CVs) are close binary systems with a white dwarf primary accreting matter supplied by a late type main-sequence secondary star via an accretion disc or along magnetic field lines of the white dwarf. Magnetic CVs, where the white dwarf has a sufficiently strong magnetic field to affect the accretion trajectory, form two distinctive subclasses: the high-field polars, and the low-field intermediate polars (IPs). These subclasses are characterized by well-defined observational properties (Cropper 1990; Patterson 1994; Warner 1995). The polars are usually soft X-ray emitters and have near synchronously rotating WD, the IPs are harder X-ray sources and show a second periodicity due to the asynchronously rotating WD. In some cases a third period is observable, which is interpreted as the beat period between orbital and spin periods. Besides differences in the flux distribution and variability, the orbital period distribution of the various subclasses of CVs were also noticed to be different (Kolb 1995). Polars tend to cluster below the period gap (2 h $< P_{orb}<$ 3 h), while IPs are preferentially above the gap. Non-magnetic CVs are distributed almost equally. All subclasses however show deficiency of systems in the period gap and a short-period cutoff at $\\rm P_{min}=80$ min (the minimum period). The statistically significant properties of the period distribution are presumed to have an evolutionary origin (Verbunt \\& Zwaan 1981, Verbunt 1984, King 1988, Kolb and Ritter 1992). The rapidly increasing number of magnetic CVs discovered from the ROSAT data has a significant impact on the above mentioned distribution and its consequences. ", "conclusions": "A new cataclysmic variable is discovered with interesting features: \\begin{enumerate} \\item The orbital period of $81\\pm 5$ min puts \\rxj\\ near the hydrogen burning period minimum where CVs experience a turning point of their evolution. Large flickering in the optical light curve and the observed spectral features of the object certainly show the presence of an accretion disc. \\item The limited search in the Sonneberg all-sky patrol plates revealed that the system undergoes outburst activity. Another outburst was recorded (vsnet-alert No. 1379) shortly after the object's discovery was announced through the VSNET (vsnet-chat No 662). From the plate statistics we can assume that the system has rather frequent outbursts. The amplitude of the outbursts of about 4 mag are typical for dwarf novae systems, but not as large as in SU~UMa superoutbursts or the so called TOADs (tremendous outburst amplitude dwarf novae; Howell \\etal\\ 1995). \\item There are periodic light variations with a period of 8.5 min in the light curve of the \\rxj. We observed them directly on four out of five occasions. In the fifth night a periodic signal with a side-band frequency was detected in the power spectrum. Very recently, the 8.5\\,min period was confirmed by R. Fried (vsnet-alert No 1387) from more prolonged observations. \\end{enumerate} Thus, \\rxj\\ shows mixed characteristics, making its type classification uncertain. From purely spectroscopic characteristics one may conclude that the new CV is a dwarf nova. Its short orbital period suggests that instead it may belong to the SU~UMa class or TOADs. But the repetitive detection of high-frequency pulses with a clearly fixed period indicates that it deserves a classification as an intermediate polar. This still needs to be confirmed by checking the coherency of the photometric pulses and by the detection of X-ray pulses. Intermediate polars are CVs with the primary white dwarf rotating asynchronously due to its moderate magnetic field. Column accretion onto the magnetic poles results in the emission of high-energy radiation. This radiation is reprocessed elsewhere in the system into optical light which is modulated at periods shorter than the orbital period. The optical modulation can track the spin and/or the spin/orbit beat period of the binary (see the review by Patterson 1994). The presence of X-ray emission in the quiescent state of \\rxj\\ along with the moderate He~{\\sc ii} 4686 \\AA\\ emission also argue in favor of a magnetic nature. The survey of non-magnetic CVs by van Teeseling \\etal\\ (1996) shows that the majority of X-ray emitting dwarf novae are of the SU UMa type, but they all are softer sources (HR1$\\le 0.7$) than \\rxj. There are a few long-period objects classified as non-SU UMa variables that are as hard as \\rxj. These belong to the VY~Scl, Z~Cam, and UX~UMa subclasses. We do not have any evidence which support a classification of \\rxj\\ as any of these types. Hence, since the rest of the CVs which are X-ray sources are magnetic, we conclude that \\rxj\\ is most probably magnetic. Only two IPs have been detected with EUVE and only a few dwarf novae, all of the latter during outburst. AM Herculis stars, particularly those with high magnetic fields are detected using EUVE. Thus, the non-detection of \\rxj\\ with EUVE does not prove that it is not an IP. It may indicate that \\rxj\\ could have a weak magnetic field (B $< 8$ MG), but given that only two IPs have EUVE detections and because of the rather large distance of \\rxj\\ the non-detection is not considered unusual. On the other hand, the intensity of the He{\\sc~ii} emission is not high enough to unambiguously classify it as a magnetic system. Silber (1992) set the following criteria for magnetic CVs: $20<$EW~(H$\\beta)<40\\AA$ and He{\\sc ii}/H$\\beta>0.4$. In our case, if the larger equivalent width could be attributed to a shorter orbital period, the He{\\sc ii}/H$\\beta$ ratio is definitely below this criterium ($\\approx~0.15$). The existence of outbursts and the short orbital period of the system is in some discordance with the IP classification. Most IPs cluster above the 2--3 hour period gap, while short period magnetic CVs are usually polars. However, a weak field IP will remain an IP even when it evolves towards shorter periods. In IPs, accretion outside of the Alfven radius remains in the form of a disc, while accretion inside the Alfven radius is dominated by flow along magnetic field lines. Outburst activity is uncommon since the inner part of the disc is disrupted by the magnetic field. Nevertheless, neither outburst activity nor short orbital period exclude the possibility of \\rxj\\ to be classified as an IP. In addition, the presence of a large disc in a short period magnetic CV suggests that the magnetic field is weak. Otherwise it would be a polar. For such a weak field case it may not be surprising that \\rxj\\ appears to be an IP with some properties (i.e. outbursts) similar to non-magnetic CVs, yet the evidence that it is a magnetic CV is compelling. Therefore, we offer \\rxj\\ as a candidate for the shortest period intermediate polar." }, "9803/astro-ph9803002_arXiv.txt": { "abstract": "We show that pulsar velocities may arise from anisotropic neutrino emission induced by resonant conversions of massless neutrinos in the presence of a strong magnetic field. The main ingredient is a small violation of weak universality and neither neutrino masses nor magnetic moments are required. ", "introduction": " ", "conclusions": "" }, "9803/astro-ph9803016_arXiv.txt": { "abstract": "The $^9$Be\\,{\\sc ii} $\\lambda$ 3131 \\AA\\ doublet has been observed in the solar-type stars 16 Cyg A \\& B and in the late G-type star $\\rho^1$ Cnc, to derive their beryllium abundances. 16 Cyg A \\& B show similar (solar) beryllium abundances while 16 Cyg B, which has been proposed to have a planetary companion of $\\sim 2$ $M_{\\rm Jup}$, is known to be depleted in lithium by a factor larger than 6 with respect to 16 Cyg A. Differences in their rotational histories which could induce different rates of internal mixing of material, and the ingestion of a similar planet by 16 Cyg A are discussed as potential explanations. The existence of two other solar-type stars which are candidates to harbour planetary-mass companions and which show lithium and beryllium abundances close to those of 16 Cyg A, requires a more detailed inspection of the peculiarities of the 16 Cyg system. For $\\rho^1$ Cnc, which is the coolest known object candidate to harbour a planetary-mass companion ($M > 0.85$ $M_{\\rm Jup}$), we establish a precise upper limit for its beryllium abundance, showing a strong Be depletion which constrains the available mixing mechanisms. Observations of similar stars without companions are required to asses the potential effects of the planetary companion on the observed depletion. It has been recently claimed that $\\rho^1$ Cnc appears to be a subgiant. If this were the case, the observed strong Li and Be depletions could be explained by a dilution process taking place during its post-main sequence evolution. ", "introduction": "\\label{sec1} In very recent years, several stars have been proposed to have planetary companions on the basis of measured precise radial velocity variations. This field of research is experiencing rapid development, and updated reviews of the present situation can be found in the proceedings of the workshop on {\\it Brown Dwarfs and Extrasolar Planets} edited by Rebolo et al. (1998) and in The Extrasolar Planets Encyclopaedia\\footnote{http://wwwusr.obspm.fr/planets/} by J. Schneider. Once a solar-type star has been suggested to harbour a planetary-mass companion, it is interesting to investigate any similarities with the Sun, as well as to find possible differences with respect to other single stars. Chemical abundances are among the most important parameters to be compared and, in particular, precise abundances of light elements such as lithium and beryllium (easy to destroy by $(p,\\alpha)$ nuclear reactions when the temperature reaches $\\sim 2.5\\times 10^6$ and $\\sim 3.5\\times 10^6$ K, respectively) combined with the abundances of other elements which are not so readily destroyed in stellar interiors, should help to understand how the presence of planets may affect the chemical composition of their parent stars. Gonzalez (1997, 1998) has derived the overall metallicities as well as abundances of different elements (including lithium) for a wide sample of proposed parent stars, finding that four of the known systems show a metallicity significantly higher than the solar value. A peculiar system such as 16 Cyg A \\& B, formed by twin solar-type stars of which only one has an orbiting planet (Cochran et al. 1997), is an especially suitable candidate to perform a detailed abundance study. Gonzalez (1998) found that both stars have a similar metallicity with a value slightly larger than solar, and confirmed independently a previous result of King et al. (1997a) that 16 Cyg B (the star with a suspected planet) is strongly depleted in lithium with respect to 16 Cyg A. The knowledge of their beryllium abundances is of potential value in quantifying the possible influence of a planetary companion on the mixing mechanisms operating in the stellar interior. $\\rho ^1$ Cnc is a star with spectral type G8V, and is the coolest known object which is a candidate to have a planetary companion. Following Gonzalez (1998), this star falls into the group having roughly Jupiter-mass companions with small circular orbits and very metal-rich parent stars. Dominik et al. (1998) have shown recently that the planetary system of $\\rho ^1$ Cnc also hosts a Vega-like disk of dust, evidenced by an infrared excess at 60 $\\mu$m. The star is very depleted in lithium and its beryllium abundance could be compared with existing upper limits measured in younger stars with similar effective temperatures (Garc\\'\\i a L\\'opez et al. 1995a). In this paper we derive the beryllium abundances of the 16 Cyg system and of $\\rho ^1$ Cnc by comparing observations with spectral syntheses of the $^9$Be\\,{\\sc ii} $\\lambda$ 3131 \\AA\\ doublet. We use those, together with their published lithium values, as well as with available abundances for other stars (with and without suggested planetary companions), and discuss briefly possible effects of planets on processes taking place in their structure and evolution. ", "conclusions": "\\label{sec5} Beryllium abundances have been derived for the solar-like stars 16 Cyg A \\& B and the cooler object $\\rho ^1$ Cnc, for which there are published values of their lithium abundances. 16 Cyg B and $\\rho ^1$ Cnc are candidates to be parents of extrasolar planets, and by measuring their Be abundances we aim at studying the potential dependence on the presence of planetary companions of detailed processes operating in their structure and evolution. 16 Cyg A \\& B show very similar Be abundances, which are compatible with the solar value, while the lithium abundance of 16 Cyg B is at least a factor 6 smaller than that of 16 Cyg A. Different rates of mixing of material in their interiors associated with different angular momentum histories, as well as the hypothetical ingestion of a planetary companion by 16 Cyg A are discussed as potential explanations. The existence of two other solar-like parent stars, whose Li (and Be) does not show strong depletion, i.e. whose behaviour is like 16 Cyg A, the Sun and the majority of similar stars with Li and Be abundances available, implies that the 16 Cyg system requires special observational and theoretical attention. A low upper limit has been derived for the beryllium abundance of $\\rho ^1$ Cnc. This is the first time a precise limit has been set and that such strong Be depletion has been observed in a late G-/early K-type MS star. This measurement clearly constrains the depletion predictions of the available mixing mechanisms, but requires observation of planet-free stars with similar age and spectral type to discard the potential effects of the planetary companion on the Li and Be depletions. Claims have also been made indicating that $\\rho ^1$ Cnc appears to be a subgiant. If this were the case, its strong Li and Be depletions could be explained by a dilution process taking place during its post-MS evolution." }, "9803/astro-ph9803199_arXiv.txt": { "abstract": "We present a study of the polarizing power of the dust in cold dense regions (dark clouds) compared to that of dust in the general interstellar medium (ISM). Our study uses new polarimetric, optical, and spectral classification data for 36 stars to carefully study the relation between polarization percentage ($p$) and extinction ($A_V$) in the Taurus dark cloud complex. We find two trends in our $p-A_V$ study: (1) stars background to the warm ISM show an increase in $p$ with $A_V$; and (2) the percentage of polarization of stars background to cold dark clouds does not increase with extinction. {\\it We detect a break in the $p-A_V$ relation at an extinction $1.3 \\pm 0.2$ mag, which we expect corresponds to a set of conditions where the polarizing power of the dust associated with the Taurus dark clouds drops precipitously. This breakpoint places important restrictions on the use of polarimetry in studying interstellar magnetic fields.} ", "introduction": "The polarization of background starlight has been used for nearly half a century to probe the magnetic field direction in the interstellar medium (ISM). The observed polarization is believed to be caused by dichroic extinction of background starlight passing through concentrations of aligned elongated dust grains along the line-of-sight. Although there is no general consensus on which is the dominant grain alignment mechanism (Lazarian, Goodman \\& Myers 1997), it is generally believed that the shortest axis of the ``typical'' elongated grain tends to be become aligned to the local magnetic field. For this orientation, the observed polarization vector is parallel to the plane-of-the-sky projection of a line-of-sight-averaged magnetic field (Davis \\& Greenstein 1951). The line-of-sight averaging inherent in background starlight polarimetric observations can make interpretation of the polarization produced by different field orientations and/or several independent dust clouds very complicated. Nonetheless, it was thought that if lines of sight with just one localized very dusty region (such as a dark cloud) between us and a background star could be found, surely the polarization would reveal the field associated with that dusty region. However, recent studies in the Taurus region (Goodman et al. 1992; Gerakines et al. 1995) and other parts of the sky (Creese et al. 1995; Goodman et al. 1995) have uncovered substantial evidence to show that dust inside cold dark clouds has lower polarizing power than dust in the general warm ISM. This means that the polarization of the light from a background star is a non-uniformly {\\it weighted} line-of-sight average of the projected plane-of-the-sky field, and that grains in cold dark clouds are systematically down-weighted. The ultimate implication of this down-weighting is that above some (column?) density threshold, the polarization of background starlight gives no information about the magnetic field in dark clouds. It is the goal of this Letter to find and physically describe this threshold. The evidence that grains in cold dark clouds are inefficient polarizers of background starlight is multi-faceted. Eight years ago, Goodman et al. (1990) found that the smooth large-scale patterns apparent in polarization maps of dark cloud complexes (e.g. Vrba et al. 1976; Vrba et al. 1981; Moneti et al. 1984; Whittet et al. 1994) do not systematically change in response to the large density enhancements represented by the dark clouds. After this realization, it was hypothesized that perhaps optical polarimetry was incapable of seeing field changes which might occur only in the high-density, optically opaque, interiors of dark clouds. So, near-infrared polarimetry, which can probe the optically opaque cloud interiors was undertaken. The near-infrared observations showed that the mean direction and dispersion of the polarization vectors are virtually {\\it identical} in the cloud interiors and their peripheries (Goodman et al. 1992; 1995). Thus, the presence of cold dark clouds still appeared to have no geometric effect on the polarization maps, implying either that: 1) the field is truly unaffected by the cloud; or 2) that background starlight polarimetry is somehow insensitive to the field in dark clouds. Polarization-extinction relations provide the best discriminant between these hypotheses. For grains of constant polarization efficiency, $p$ should rise with $A_V$. Using the near-infrared observations, Goodman et al. (1992, 1995) find that the {\\it percentage of polarization does not rise with extinction} in cold dark clouds. The simplest interpretation\\footnote{Note that increased field tangling inside dark clouds cannot explain the near-infrared polarimetric observations for two reasons. 1.) The dispersion in the distribution of position angle does not increase in the cloud interior (near-IR observations) relative to the periphery (optical observations). And, 2.) while it is true that the slope of a $p-A_V$ relation will diminish due to field tangling, it will remain positive even for highly tangled fields if all grains polarize equally well (see Jones 1989; Jones et al. 1992).} of this result is that dust in dark clouds adds plenty to the observed extinction, but has very little ``polarizing power\" and so adds only a very small fraction to the observed net polarization. A number of factors, including poor grain alignment, grain growth, and/or changes in grain shape or composition, could be responsible for the low polarization efficiency exhibited by dust grains in cold dark clouds (see Goodman et al. 1995). Regardless of which factor(s) cause(s) the low polarization efficiency exhibited by by dust in dark clouds, the fact is that background starlight polarimetry does not reliably reveal the magnetic field {\\it in} dark clouds. Based on the near-infrared studies, we expect that the fraction of grains with high polarization efficiency is relatively constant in the lower-density warm ISM, but drops precipitously in dark clouds. Therefore, we hypothesize that a breakpoint in the $p-A_V$ relationship might exist at the dark clouds' ``edges\", which previous studies in the near-IR (Goodman et al. 1992; 1995; Gerakines et al. 1995) could not detect, due to their inability to measure low $A_V$'s accurately enough. In this Letter, we present our attempt to carefully study the $p-A_{V}$ relation near dark clouds, and thus offer a set of guidelines as to where the polarization maps can be taken as faithful representations of the magnetic field projected onto the plane of the sky, and where they cannot. ", "conclusions": "The breakpoint in the $p-A_{V}$ relation places important restrictions on the use of polarimetry in studying interstellar magnetic fields. Since the polarization efficiency of the dust inside dark clouds is very low, most of the polarization observed for lines of sight which pass through these extinction peaks is not due to the dark cloud; it is due to dust background and foreground to the cloud. Hence, one should not use background starlight polarimetry to map magnetic fields inside dark clouds. With the results of this study we can quantify the word ``inside.\" In regions like Taurus, {\\it it is safe to interpret the polarization of background starlight as a representation of the plane-of-the-sky projected magnetic field up to the 1.3 $\\pm$ 0.2 mag ``edge'' of the dark cloud}. In other words the linear relation between $p$ and $A_V$ that exists in the low-density ISM breaks down for stars background to the $\\simgreat 1.3$ mag of extinction produced by a dense localized dusty region (i.e., dark cloud). After this edge polarization no longer rises with extinction, and thus cannot reveal the field structure in the dense region. We restate that this proscription only applies for stars background to cold dark clouds, as stars background to the warm ISM have not been shown to exhibit such behavior." }, "9803/astro-ph9803150_arXiv.txt": { "abstract": "We investigate the effect of gravitational lensing by matter distribution in the universe on the cosmic microwave background (CMB) polarization power spectra and temperature-polarization cross-correlation spectrum. As in the case of temperature spectrum gravitational lensing leads to smoothing of narrow features and enhancement of power on the damping tail of the power spectrum. Because acoustic peaks in polarization spectra are narrower than in the temperature spectrum the smoothing effect is significantly larger and can reach up to 10\\% for $l<1000$ and even more above that. A qualitatively new feature is the generation of $B$ type polarization even when only $E$ is intrinsically present, such as in the case of pure scalar perturbations. This may be directly observed with Planck and other future small scale polarization experiments. The gravitational lensing effect is incorporated in the new version (2.4) of CMBFAST code. ", "introduction": "Over the next few years a number of ground based, balloon and satellite experiments will measure CMB sky with an unprecedented accuracy and detail. The promise of a one percent precision on the measured power spectrum of CMB anisotropies requires a similar accuracy in the theoretical predictions, if we are to exploit all the information present in the data. The rewards will be rich: among other things this will allow an accurate determination of a number of cosmological parameters and testing of current structure formation theories \\cite{parameters}. In principle such a program is possible, since the anisotropies were produced when the universe was still in the linear regime, which makes the calculations of model predictions very accurate. In practice there are a number of important effects that need to be included if this goal is to be realized. One of the most important among these is the gravitational lensing effect. As photons propagate through the universe from their last scattering to our detectors they are randomly deflected by the gravitational force exerted upon them by the inhomogeneous mass distribution. Previous work has shown that gravitational lensing has an effect on the temperature anisotropy power spectrum which is not insignificant \\cite{others,uroslens}. The random deflections smear out the sharp features in the correlation function or power spectrum, leading to a suppression of acoustic oscillations. Gravitational lensing can also enhance power on the damping tail, causing it to decay less rapidly than predicted on very small angular scales \\cite{bentonsilk}. Gravitational lensing effect on the temperature anisotropies has been discussed several times in the literature and the formalism to calculate it using the evolution of density power spectrum both in linear and nonlinear regime has been presented in \\cite{uroslens}. In this paper we extend this calculation to the two linear polarization power spectra and to the cross-correlation spectrum between temperature and polarization. Because acoustic oscillations are narrower for polarization spectra than for temperature, one expects gravitational lensing effect to be more significant in the former and indeed our results confirm this. In addition, a qualitatively new effect is the mixing between $E$ and $B$ types of polarization, which changes the pattern of polarization. The outline of the paper is the following. In \\S \\ref{formalism} we develop the formalism: this section contains all the main analytic expressions needed for a numerical implementation of the effect. These have been numerically implemented in the new version of CMBFAST package (version 2.4) and require only a marginal increase in the CPU time for their evaluations. In \\S \\ref{estimate} we compute the effect for a typical cosmological model and address the question of direct observability of the effect. We present the conclusions in \\S \\ref{conclusions}. ", "conclusions": "" }, "9803/astro-ph9803293_arXiv.txt": { "abstract": "Optical multi-slit spectra have been obtained for 47 globular clusters surrounding the brightest Virgo elliptical NGC~4472 (M49). Including data from the literature, we analyze velocities for a total sample of 57 clusters and present the first tentative evidence for kinematic differences between the red and blue cluster populations which make up the bimodal colour distribution of this galaxy. The redder clusters are more centrally concentrated and have a velocity dispersion of 240 kms$^{-1}$ compared with 320 kms$^{-1}$ for the blue clusters. The origin of this difference appears to be a larger component of systematic rotation in the blue cluster system. The larger rotation in the more extended blue cluster system is indicative of efficient angular momentum transport, as provided by galaxy mergers. Masses estimated from the globular cluster velocities are consistent with the mass distribution estimated from X-ray data, and indicate that the M/L$_B$ rises to $50$~M/L$_\\odot$ at 2.5 R$_e$. ", "introduction": "The study of extragalactic globular cluster systems can provide important clues to the formation history of their host galaxies. This is particularly true for elliptical galaxies for which there are two currently popular paradigms. One paradigm is the standard monolithic collapse model in which elliptical galaxies form in a single burst of star formation at high redshift (e.g. Arimoto \\& Yohii 1987). In contrast, hierarchical structure formation theories predict that spheroidal galaxies form continuously through a sequence of galaxy mergers (e.g. Cole \\etal 1994; Kauffmann 1996). Ashman \\& Zepf (1992) explored the properties of globular clusters in models in which elliptical galaxies are the products of the mergers of spiral galaxies, and showed that the greater specific frequency of globular clusters around ellipticals relative to spirals could be explained if globular clusters form during the mergers. They also predicted that elliptical galaxies formed by mergers will have two or more populations of globular clusters - a metal-poor population associated with the progenitor spirals, and a metal-rich population formed during the merger. In contrast, monolithic collapse models naturally produce unimodal metallicity distributions. The discovery that the globular cluster systems of several elliptical galaxies have bimodal colour (and by implication metallicity) distributions (Zepf \\& Ashman 1993; Whitmore \\etal 1995; Geisler \\etal 1996) provides strong support for the merger model. Geisler \\etal (1996) and Lee et al. (1998) also show that the red (metal-rich) cluster population is more centrally concentrated than the blue (metal-poor) population, as predicted by Ashman \\& Zepf (1992). Recently, an alternative view has been presented by Forbes et al.\\ (1997), who suggest that the bimodal color distributions may not be due to mergers, but to a multi-phase single collapse. Although the primary physical mechanism known to produce distinct formation episodes is mergers, it is important to attempt to distinguish between these competing models for the formation of globular cluster systems and their host elliptical galaxies. The kinematics of globular cluster systems may offer such a test of these models. In the multi-phase collapse picture, angular momentum conservation requires that the spatially concentrated metal-rich population rotates more rapidly than the extended metal-poor population. In contrast, simulations of merger models indicate that mergers typically provide an efficient means of angular momentum tranfer, and that the central regions have specific angular momentum that is lower than the outer regions (Hernquist 1993; Heyl, Hernquist and Spergel 1996). Studies of the kinematics of globular cluster systems therefore provide important constraints on the formation history of elliptical galaxies. They also provide useful probes of the mass distribution of elliptical galaxies at radii larger than can be reached by studies of the integrated light. The extended nature of globular cluster systems allows the dynamical mass determined from their velocities to be compared at similar radii to masses determined through studies of the hot X-ray gas. A recent example is the study of the M87 globular cluster system by Cohen \\& Ryzhov (1997), who find that a rising mass-to-light ratio is required out to radii of $\\sim 3 R_e$, in agreement with X-ray mass determinations. However, M87 occupies a privileged position at the center of the Virgo cluster, so it is critical to test whether the rising mass-to-light ratio, and the agreement with X-ray masses, is true for more typical cluster elliptical galaxies. In this paper we present a spectroscopic study of the globular cluster system of the elliptical galaxy NGC~4472 (M49). This is the brightest elliptical galaxy in the Virgo cluster and has been the subject of a detailed photometric study by Geisler \\etal (1996). The only previously published spectroscopic data for the NGC~4472 globular cluster population is by Mould \\etal (1990) who presented velocities and line strengths for 26 clusters. The outline of our paper is as follows: Section 2 discusses the sample selection together with our observations and data reduction; Section 3 discusses the kinematics of the metal-rich and metal-poor populations in the context of the merger model, and analyses the implications for the overall M/L ratio in NGC~4472. Finally, we present our conclusions in Section 4. ", "conclusions": "We have made a detailed spectroscopic study of the globular cluster system of NGC~4472, and have more than doubled the number of confirmed clusters to 57. Whilst this remains a statistically small sample, the data show several interesting properties when combined with the accurate colour/metallicity data from Geisler \\etal (1996). When the complete sample is divided into a metal-rich (47\\%) and a metal-poor (53\\%) subset on the basis of their bimodal colour histogram, the metal-poor subset appears to have a broader distribution of velocities. We have investigated this further, and conclude that the most likely cause is a higher mean level of rotation in the metal-poor cluster system, which is consistent with that of the underlying stellar halo in amplitude and position angle (but with a much higher specific angular momentum). The metal-rich clusters on the other hand show only weak evidence for any rotation, and about an axis which is tilted $\\sim 50^o$ from that of the other components. These results are qualitatively in agreement with the predictions of a model in which the metal-rich clusters are formed during the merger of two massive gas-rich galaxies, each with its own old metal-poor cluster population. The cluster system of NGC~4472 forms a dynamically hotter population than the stellar halo, but is consistent with being in dynamical equlibrium with the halo potential defined by the hot X-ray emitting plasma, and supports the presence of a dark $\\sim 10^{12}{\\rm M}_\\odot$ halo in this giant elliptical galaxy." }, "9803/hep-ph9803309_arXiv.txt": { "abstract": "The two possible Mikheyev-Smirnov-Wolfenstein (MSW) solutions of the solar neutrino problem (one at small and the other at large mixing angle), up to now tested mainly through absolute neutrino flux measurements, require flux-independent tests both for a decisive confirmation and for their discrimination. To this end, we perform a joint analysis of various flux-independent observables that can be measured at the SuperKamiokande and Sudbury Neutrino Observatory (SNO) experiments. In particular, we analyze the recent data collected at SuperKamiokande after 374 days of operation, work out the corresponding predictions for SNO, and study the interplay between SuperKamiokande and SNO observables. It is shown how, by using only flux-independent observables from SuperKamiokande and SNO, one can discriminate between the two MSW solutions and separate them from the no oscillation case. ", "introduction": " ", "conclusions": "" }, "9803/astro-ph9803250_arXiv.txt": { "abstract": "This study shows one important effect of preexistent cosmic microwave background temperature fluctuations on the determination of the Hubble constant through Sunyaev-Zel'dovich effect of clusters of galaxies, especially when coupled with the gravitational lensing effect by the same clusters. The effect results in a broad distribution of the apparent Hubble constant. The combination of this effect with other systematic effects such as the Loeb-Refregier Effect seems to provide an explanation for the observationally derived values of the Hubble constant currently available based on the Sunyaev-Zel'dovich effect, if the true value of the Hubble constant is $60-80~$km/s/Mpc. It thus becomes possible that the values of the Hubble constant measured by other techniques which generally give a value around $60-80~$km/s/Mpc be reconciled with the SZ effect determined values of the Hubble constant, where are systematically lower than others and have a broad distribution. ", "introduction": "The Sunyaev-Zel'dovich (SZ) effect of a cluster of galaxies on the cosmic microwave background (CMB) photons can be used to determine the distance to the cluster hence the Hubble constant ($H_0$), when analysed in conjunction with X-ray observations of the cluster (Cavaliere, Danese, \\& De Zotti 1977; Gunn 1978; Silk \\& White 1978; Birkinshaw 1979). For an excellent recent review on this subject and other SZ related topics, see Rephaeli (1995 and references therein). The accuracy of the Hubble constant determination depends upon the accuracy of several assumptions involving both sets of observations (radio and X-ray). Perhaps among the most important are the assumptions of sphericity, isothermality of clusters of galaxies (e.g., Inagaki \\etal 1995). In this {\\it Letter} we point out a completely separate effect on the determination of the Hubble constant due to preexistent, small-amplitude CMB temperature fluctuations before the photons undergo the SZ effect through a cluster. The effect is significantly amplified by the gravitational lensing of the CMB photons by the cluster, because the SZ observational beam size is typically comparable to Einstein radius of the source-lens system. This effect, when coupled with some systematic effects such as the one proposed by Loeb \\& Refregier (1997) due to the systematic over-removal of background point radio sources in the beam, may provide an explanation for the observed distribution of $H_0$ determined by SZ effect. ", "conclusions": "We show that the background CMB fluctuations, especially when they are coupled with the gravitational lensing effect by clusters of galaxies, have one important effect on the determination of the Hubble constant through Sunyaev-Zel'dovich effect of the clusters (for the adopted set of characteristic numbers for the cluster, which seem fairly realistic compared to those of real clusters of interest): a broad distribution of the apparent Hubble constant is produced with a FWHM about $30\\%$ of the apparent mean value. The combination of this effect with other systematic effects such as the Loeb-Refregier Effect seems to provide a reasonable explanation for the observationally derived values of the Hubble constant currently available, if the true value of the Hubble constant is $\\sim 65~$km/s/Mpc. Thus, it becomes possible that the values of $H_0$ measured by other techniques which generally give a value around $60-80$km/s/Mpc [e.g., $73\\pm 10~$km/s/Mpc ($1\\sigma$) from Freedman, Madore, \\& Kennicutt 1997 based on HST observations of Cepheids; $64\\pm 6~$km/s/Mpc ($1\\sigma$) from Riess, Press, \\& Kirshner 1996 based on type Ia supernova multicolor light-curve shapes; $64\\pm 13~$km/s/Mpc ($95\\%$ confidence level) from Kundic \\etal 1997 based on gravitational lensing time delay measurements; $70\\pm 5~$km/s/Mpc ($1\\sigma$) from Giovanelli 1997 using I-band Tully-Fisher relation; $81\\pm 6~$km/s/Mpc ($1\\sigma$) from Tonry 1997 using surface brightness fluctuations] be reconciled with the SZ effect determined values of $H_0$. It may be possible, at least in principle, that one can use a large sample of SZ measured Hubble constant to infer the fluctuations of the CMB at the relevant scales, when the Hubble constant is independently measured to high accuracy by methods such as that using detached eclipsing binaries (Paczynski 1997)." }, "9803/astro-ph9803299_arXiv.txt": { "abstract": "{A mechanism of ultra-high energy cosmic ray acceleration in extragalactic radio sources, at the interface between the {\\it relativistic} jet and the surrounding medium, is discussed as a supplement to the shock acceleration in `hot spots'. Due to crossing the tangential discontinuity of the velocity the particle can gain an amount of energy comparable to the energy gain at the shock crossing. However, the spectrum of particles accelerated at the jet side boundary is expected to be much flatter than the one formed at the shock. Due to this fact, particles accelerated at the boundary can dominate the overall spectrum at highest energies. In conditions characteristic to extragalactic jets' terminal shocks, the mechanism naturally provides the particles with $E \\sim 10^{20}$ eV and complies with the efficiency requirements. The spectrum formation near the cut-off energy due to action of both the shock acceleration and the tangential discontinuity acceleration is modelled with the Monte Carlo particle simulations. It confirms that the upper energy limit can surpass the shock acceleration estimate.} ", "introduction": "Jet-like outflows are observed in a number of astrophysical environments, starting from young stars embedded in their parent molecular clouds, up to active extragalactic objects. In the later case, a number of interesting observational phenomena are noted over orders of magnitude of linear sizes. In particular, at the smallest mili-arc-second scales, one often observes relativistic jet velocities, with flow Lorentz factors $\\gamma_u$ reaching the values above $10$ (cf. Ghisellini et al. 1996). At larger scales, velocity measurements are more difficult, but without entrainment of large amount of matter near the active galactic nuclear source the jet flow velocity must be also relativistic. A possible loading of a jet with matter is expected to be appended by a substantial amount of turbulence (Henriksen 1987) and related jet kinetic energy dissipation. However, in the FR~II radio sources, there are often observed jets efficiently transporting energy to the far-away hot spots and any jet breaking mechanism can not act too effectively near the central core. Also, the existing hydrodynamical simulations of relativistic jets show for possibility of extended stable jet structures (Marti et al. 1995, 1997; G\\'omez et al. 1995). Another argument suggesting the relativistic jet speed at all scales, may be based on the visible asymmetry of jets with respect to the nuclear source, if one believes the effect is caused by the high velocity of the essentially bi-symmetric outflow (cf. Bridle et al. 1994). Let us also note that the Meisenheimer et al. (1989) modelling of the shock acceleration process at extragalactic radio-source hot spots yields `the best-guess' jet velocities in the range ($0.1$, $0.6$) The relativistic movement of the jet leads to shock wave formation in places where an obstacle or perturbation of the flow creates a sudden velocity jump. For jets loaded with a cold plasma the highly oblique conical shocks are formed within the jet tube. These shocks can have a non-relativistic character, involving the velocity jump perpendicular to the shock surface much smaller than the overall jet velocity $U \\sim c$. They lead to a limited kinetic energy dissipation and are usually claimed to be responsible for forming the so called `knots' along the jet. A much more powerful shock is formed at the final working surface of the jet. There, a substantial fraction of the jet energy is transferred into heating the jet's plasma, generating strong turbulence, boosting magnetic fields within the turbulent volume, and finally accelerating electrons and nuclei to cosmic ray energies. Rachen \\& Biermann (1993) considered the process of particle acceleration to ultra-high energies (UHE) at such shocks. They show that given the favourable conditions the UHE particles up to $\\sim 10^{20}$ eV can be formed. Then Rachen et al. (1993) show that assumption of UHE particle acceleration in extragalactic powerful radio sources is compatible with the current measurements of cosmic ray abundances and spectra at energies above $10^{17}$ eV. Additionally, the arrival directions of cosmic ray particles observed above $10$ EeV are correlated with the local galactic supercluster structure (Stanev et al. 1995; see, also, Medina Tanco et al. 1996, Sigl 1996, Sigl et al. 1995, 1996, Hayashida et al. 1996, Elbert \\& Sommers 1995, Geddes et al. 1996 and Medina Tanco 1998). An alternative model involving the several-Mpc-scale non-relativistic shocks in galaxy clusters is proposed by Kang et al. (1996; see also Kang et al. 1997). As noted by us (Ostrowski 1990; henceforth Paper~I) a tangential discontinuity of the velocity field can also provide an efficient cosmic ray acceleration site if the considered velocity difference $U$ is relativistic and the sufficient amount of turbulence on both its' sides is present. The problem was extensively discussed in the early eighties by Berezhko with collaborators (see the review by Berezhko 1990) and in the diffusive limit by Earl et al. (1988) and Jokipii et al. (1989). In the present paper we consider the process of ultra high energy cosmic ray acceleration in relativistic jets including the possibility of such boundary layer acceleration. As the considerations of Rachen \\& Biermann (1993) treat the acceleration process at relativistic shock in a somewhat simplified way (see, also Sigl et al. 1995), in the first part of the next section (section 2.1) we review this process in some detail in order to understand the inter-relations between the conditions existing near the shock, the accelerated particle spectrum and the particle's upper energy limit. Then, in section (2.2), we present a short description of the basic physical model for the considered acceleration process acting at the jet boundary. We show (section 2.3) that in the conditions characteristic for relativistic jets in extragalactic radio sources, particles with energies above $10^{20}$ eV can be produced in this process without extreme parameter fitting. The required efficiency is discussed in section (2.4). We confirm the estimates presented previously for the shock acceleration, showing that the UHE particle flux observed at the Earth can be reproduced as a result of acceleration processes in jets of nearby powerful radio sources. In section 3 we discuss the problem of the particles' spectrum. With the use of Monte Carlo simulations, we consider the action of both processes acting near the terminal shock in a relativistic jet. Modification of the spectrum due to varying boundary conditions and jet velocity is discussed for the case of ($e^-$, $p$) jets expected to occur in the powerful FRII radio sources (cf. Celotti \\& Fabian 1993). The derived particle's upper energy limits are above the shock acceleration estimates and the spectrum modification at highest energies can resemble the observed above 10 EeV `ankle' structure. A short summary and final remarks are presented in section 4. A preliminary report about this work was presented in Ostrowski (1993b, 1996). For the discussion that follows, we consider the jet propagating with the relativistic velocity, $U \\sim c$. We use $c$ = 1 units. ", "conclusions": "" }, "9803/astro-ph9803066_arXiv.txt": { "abstract": "We present new Keck observations of giant arcs in the cluster Abell 2390. High resolution two-dimensional spectra of two arcs show metal lines at $z=4.040\\pm 0.005$ with Ly$\\alpha$ emission spatially separated from the stellar continuum. In addition, spectroscopy along the notorious `straight' arc reveals two unrelated galaxies, at z=0.913 and z=1.033, with absorption lines of MgII and FeII at z=0.913 seen against the more distant object, indicating the presence of a large gaseous halo. ", "introduction": "New high resolution spectra are presented for the high-redshift $z=4.04$ lensed system behind A2390. Figure 1 shows a section of the spectrum aligned along the long axes of the two main arcs. This clearly shows that the Ly$\\alpha$ emission is spatially-separated from the continuum light and redshifted with respect to the interstellar lines, indicating an outward flow of enriched gas thought to be typical of starburst galaxies (Lequeux et al. 1995). All stellar and interstellar features are common to both spectra, confirming these arcs are images of one single highly-magnified galaxy. Note the Ly$\\alpha$ absorption is seen only in the southern portions of both arcs, coincident with the stellar continuum and with no associated Ly$\\alpha$ emission, indicating absorption of this line where the HI column is high. A lens model for this system is discussed in Frye \\& Broadhurst (1998). Keck infrared observations were also taken to study the stellar populations of this galaxy, by comparing flux ratios on opposite sides of the 4000 \\AA \\ break (Bunker, et al. these proceedings). ", "conclusions": "" }, "9803/astro-ph9803316_arXiv.txt": { "abstract": "We have carried out K-band speckle observations of a sample of 114 X-ray selected weak-line T\\,Tauri stars in the nearby Scorpius-Centaurus OB association. We find that for binary T\\,Tauri stars closely associated to the early type stars in Upper Scorpius, the youngest subgroup of the OB association, the peak in the distribution of binary separations is at 90 A.U. For binary T\\,Tauri stars located in the direction of an older subgroup, but not closely associated to early type stars, the peak in the distribution is at 215 A.U. A Kolmogorov-Smirnov test indicates that the two binary populations do not result from the same distribution at a significance level of 98\\%. Apparently, the same physical conditions which facilitate the formation of massive stars also facilitate the formation of closer binaries among low-mass stars, whereas physical conditions unfavorable for the formation of massive stars lead to the formation of wider binaries among low-mass stars. The outcome of the binary formation process might be related to the internal turbulence and the angular momentum of molecular cloud cores, magnetic field, the initial temperature within a cloud, or - most likely - a combination of all of these. We conclude that the distribution of binary separations is not a universal quantity, and that the broad distribution of binary separations observed among main-sequence stars can be explained by a superposition of more peaked binary distributions resulting from various star forming environments. The overall binary frequency among pre-main-sequence stars in individual star forming regions is not necessarily higher than among main-sequence stars. ", "introduction": "Taurus-Auriga is the star forming region which has been most thoroughly surveyed for pre-main-sequence binary and multiple systems (see Mathieu 1994 for a review). For separations from 15 A.U.\\ to 2000 A.U., the binary frequency among T\\,Tauri stars in Taurus is 1.9 times as high as among nearby main-sequence stars (K\\\"ohler \\& Leinert 1998). Extrapolating over the whole range of separations yields a binary frequency of 100\\%, i.\\,e., each T\\,Tauri star in Taurus should be member of a binary or multiple systems. This apparent overabundance of binaries among pre-main-sequence stars is puzzling. One possible explanation is a decrease in the binary frequency as a function of the age of a stellar population (Patience et al.\\ 1998). However, a T\\,association like Taurus might not be the typical birthplace for low-mass stars, as up to 80\\% of all low-mass stars could originate in OB associations (Miller \\& Scalo 1978; see also Zinnecker et al.\\ 1992). Scorpius-Centaurus is the most nearby OB association at a distance of about 145 parsec (de Zeeuw et al.\\ 1998). It consists of three subgroups (cf.\\ Figure 1) with ages ranging from 5 to 13 Myr (de Geus et al.\\ 1989). Upper Centaurus-Lupus (UCL) is the oldest subgroup of the association. Star formation started here 13 Myr ago and subsequently progressed throughout the parental giant molecular cloud (e.g.\\ Blaauw 1991 and references therein). Based on observations with the EINSTEIN X-ray satellite, Walter et al.\\ (1994) identified 28 weak-line T\\,Tauri stars (WTTS) in Upper Scorpius (US), the youngest subgroup. 10 of these have been surveyed by Ghez et al.\\ (1993) for binary or multiple systems, and three binaries have been detected. The EINSTEIN fields covered only a small fraction of US (Fig.\\ 2) and the 28 WTTS did not allow for a statistical meaningful study of binary frequencies and separations. A search for visual binary stars among 74 ROSAT selected WTTS and post-T\\,Tauri stars in US (Kunkel et al., in prep) was carried by Brandner et al.\\ (1996). This survey was sensitive to binary separations down to 0\\farcs8, and revealed a rather high binary frequency in the region located between US and UCL (`US-B') and an apparent absence of wide visual binary stars in the center of US (`US-A'). In order to identify closer binary systems and to get a better statistics on possible spatial variations of binary star properties, we have carried out a speckle survey of 114 WTTS in US based on the lists by Walter et al.\\ (1994) and Kunkel et al.\\ (in prep). \\begin{figure*}[htb] \\centerline{\\plotfiddle{fig1.ps}{11cm}{0}{90}{90}{-520}{-310}} \\figcaption{The spatial distribution of 532 proper motion members of the Scorpius-Centaurus association based on HIPPARCOS measurements (adapted from de Zeeuw et al.\\ 1998). The boundaries between the subgroups Upper Scorpius (US), Upper Centaurus Lupus (UCL), and Lower Centaurus Crux (LCC) are indicated by dotted lines. Star formation has progressed from the oldest subgroup UCL towards the younger subgroups US and LCC. The two adjacent fields of our survey for binary T\\,Tauri stars, centered on US (US-A, solid lines) and at the interface between US and UCL (US-B, dashed lines) are outlined. \\label{fig1}} \\end{figure*} ", "conclusions": "We have shown that the distributions of binary separations among weak-line T\\,Tauri stars in two adjacent fields (`US-A' and `US-B') in the Scorpius-Centaurus OB association are clearly distinct from each other and considerably more peaked than the (broad) distribution of binary separations observed among main-sequence field stars. In US-A, the WTTS are closely associated with B type stars, whereas in US-B only a few early type stars are present. We conclude that the same physical conditions which facilitate the formation of massive stars also facilitate the formation of closer binaries among low-mass stars, whereas physical conditions unfavorable for the formation of massive stars lead to the formation of wider binaries among low-mass stars. The outcome of the binary formation process might be determined by the critical density at which the magnetic field support breaks down, the internal turbulence and the angular momentum of molecular cloud cores, the initial temperature within a cloud, or - most likely - a combination of all of these. We further conclude that the distribution of binary separations is not a universal quantity. Instead, both the peak and the width of the distribution might vary from one star forming region to the next. The broad distribution of binary separations observed among main-sequence field stars can be understood as a superposition of binary populations originating in various star forming environments with very distinct peaks in the distribution of binary separations. The apparent overabundance of binaries among T\\,Tauri stars in the Taurus-Auriga T\\,association might be explained by the fact that the distribution of binary separations there is strongly peaked towards $\\approx$ 30\\,A.U. Extrapolating from this very pronounced peak over the whole range of possible binary separations then leads to an erroneously high estimate of the overall binary frequency." }, "9803/astro-ph9803120_arXiv.txt": { "abstract": "We investigate the non-gaussian properties of cosmic-string-seeded linear density perturbations with cold and hot dark matter backgrounds, using high-resolution numerical simulations. We compute the one-point probability density function of the resulting density field, its skewness, kurtosis, and genus curves for different smoothing scales. A semi-analytic model is then invoked to provide a physical interpretation of our results. We conclude that on scales smaller than $1.5{(\\Omega h^2)}^{-1}$Mpc, perturbations seeded by cosmic strings are very non-gaussian. These scales may still be in a linear or mildly non-linear regime in an open or $\\Lambda$-universe with $\\Gamma=\\Omega h \\lsim 0.2$. ", "introduction": "\\label{intro} At present, the two main candidates for the origin of cosmic structure are inflation and topological defects (for a review, see Vilenkin \\& Shellard 1994). Although both scenarios may produce a power spectrum of density perturbations consistent with observations, they have very different predictions regarding the statistical properties of the density field. While most inflationary models produce gaussian random-phase initial conditions, defect models produce non-gaussian perturbations particularly on small scales. New results from cosmic-string-seeded structure formation using high-resolution simulations (\\markcite{ASWAs,ASWAl}Avelino {\\it et al.}~1997, 1998) were encouraging for models with $\\Gamma = \\Omega h = 0.1$--$0.2$ (see also Battye {\\it et al.}~1997); both the mass fluctuation amplitude at $8 h^{-1}$Mpc, $\\sigma_8$, and the power spectrum shape of cosmic-string-induced cold dark matter fluctuations, ${\\cal P}(k)$, were consistent within uncertainties with observational data (Peacock \\& Dodds 1994; Viana \\& Liddle 1996). However, because cosmic strings induce non-gaussian density perturbations on small scales, the power spectrum alone is insufficient to describe all the statistical properties of such a density field. This is even more important in open or $\\Lambda$-models because in those models the characteristic scales of the density field are shifted to larger scales relative to a flat model with $\\Lambda=0$. In this Letter we investigate the non-gaussian properties of the linear density field induced by cosmic strings using higher-order statistics such as the skewness and the kurtosis of a one-point probability density function (PDF), as well as genus statistics. The non-gaussian properties we reveal provide a significant observational signature for cosmic string-seeded structure formation models on small length-scales. We note that previous analytic work has investigated the string-induced velocity field on scales above several $h^{-1}$Mpc, which was inferred to be gaussian (Vachaspati 1992; Moessner 1995), and that some of the features we study here were also observed in global topological defect models, notably for textures (Park, Spergel, \\& Turok 1991). Past work on genus statistics in the context of topological defects was made using toy models which incorporated some important features of the models in question (Brandenberger, Kaplan \\& Ramsey 1993; Albrecht \\& Robinson 1995; Avelino 1997). Our first step in the present analysis was to perform high-resolution numerical simu\\-lations of cosmic string networks in an expanding universe (Allen \\& Shellard 1990) from which we subsequently computed the causally-sourced density field with either a cold or hot dark matter (CDM or HDM) background. The cosmic string simulations had a dynamic range extending from before the radiation-matter transition at $0.4 \\eta_{\\rm eq}$ through to deep into the matter era $8.4 \\eta_{\\rm eq}$, where $\\eta_{\\rm eq}$ is the conformal time at radiation-matter density equality. The structure formation simulation boxes contained $256^3$ grid-points and their physical volume was in the range $(4$--$100 h^{-1}$Mpc$)^3$. A much more detailed description of these methods is given by Avelino {\\it et al.}~(1997, 1998). ", "conclusions": "We conclude that on length scales smaller than\\ $1.5 {(\\Omega h^2)}^{-1}{\\rm Mpc}$ perturbations seeded by cos\\-mic strings are very non-gaussian, especially in the context of a CDM model. In an open or $\\Lambda$-universe with $\\Gamma=\\Omega h \\sim 0.15$, this scale will be shifted to $10 h^{-1}$Mpc, which may still be in the linear or mildly non-linear regime, thus potentially providing a strong empirical test for cosmic string models. It has been suggested that such non-gaussianity may imply that it is difficult in string models to deduce the parameter $\\beta=\\Omega_0^{0.6}/b$ from observations of density and velocity fields (van de Bruck, 1997). However, our results indicate otherwise on large scales because we find that string perturbations are very similar to gaussian-random phase fluctuations, when smoothed on sufficiently large scales." }, "9803/astro-ph9803193_arXiv.txt": { "abstract": "Radio observations establish the B/A magnification ratio of gravitational lens 0957+561 at about 0.75. Yet, for more than 15 years, the {\\it optical} magnfication ratio has been between 0.9 and 1.12. The accepted explanation is microlensing of the optical source. However, this explanation is mildly discordant with (i) the relative constancy of the optical ratio, and (ii) recent data indicating possible non-achromaticity in the ratio. To study these issues, we develop a statistical formalism for separately measuring, in a unified manner, the magnification ratio of the {\\it fluctuating} and {\\it constant} parts of the light curve. Applying the formalism to the published data of Kundi\\'c et al. (1997), we find that the magnification ratios of fluctuating parts in both the g and r colors agrees with the magnification ratio of the constant part in g-band, and tends to disagree with the r-band value. One explanation could be about $0.1$ mag of consistently unsubtracted r light from the lensing galaxy G1, which seems unlikely. Another could be that 0957+561 is approaching a caustic in the microlensing pattern. ", "introduction": "\\label{sect:intro} The gravitational lens system 0957+561 has by now been observed at optical and radio wavelengths for nearly twenty years (Walsh, Carswell, and Weymann 1979; Porcas et al. 1979). Radio studies have definitively established that the B/A magnification ratio of the lens, measured at the core of the radio images (which lies at the location of the optical point source images), is close to 0.75; some recent measurements are $0.75\\pm 0.02$ (Garrett et al. 1994), $0.752\\pm 0.028$ (Conner et al., 1992). Note that while there is some controversy about the radio magnification ratio at the location of the radio jet, as opposed to the core (see, e.g., Garrett et al. 1994), only the core value interests us here. Because the variability of the quasar in the optical is larger than in the radio, measurement of the B/A magnification ratio in the optical requires that the light curves be shifted by the correct time delay $\\tau$ before the ratio is taken. Thus, the earliest determinations of B/A were incorrect. For example, Young et al. (1980) obtained a ratio of 0.76, comfortingly -- yet erroneously -- close to the radio magnification. However, at least from Vanderriest et al. (1989) on, who used a value for $\\tau$ quite close to definitive recent determinations (Kundi\\'c et al. 1997), it has been clear that the B/A magnification ratio of the optical continuum in the B and A point sources is quite different from 0.75, and moreover has remained at least fairly constant for the full history of observation. Smoothing over observing seasons, Vanderriest et al. (1989) obtained a B/A ratio varying between about 0.9 and 1.05 over the observing seasons early-1980 through early-1986 (times referenced to A component), with a single best-fit value of 0.97. It is debatable whether the variation around the best-fit value is actual time variation of the lens magnification ratio (as distinct from time variation in the quasar luminosity, n.b.) or observational artifact. However, it does seem quite likely that the magnification ratio varied by no more than about $\\pm 8$\\% during this time. More recently, the value recently obtained by Kundi\\'c et al. (1995, 1997) for the 1995 season (A component) is $1.12\\pm 0.01$ in g-band (with the error bar, a 95\\% confidence limit, depending somewhat on the method of reduction used). So, it is quite plausible (and not contradicted by other measurements in the literature) that the optical B/A remained in the range 0.9 to 1.12 from 1980 through 1995, and possible that the variation has been considerably smaller than this range. The discrepancy between the optical and radio magnification ratios has long been understood as due to microlensing (as predicted by Chang and Refsdal 1979, and Gott 1981). The proper radius of the Einstein ring from a 0.5 $M_\\odot$ star at the lens galaxy redshift $z=0.36$, illuminated by the quasar at redshift $z=1.41$, is about $2\\times 10^{16}\\,h^{-1/2}$cm. Since the radio emission region is much larger than this scale, it averages spatially over the microlensing pattern and is magnified by the macrolens ratio of 0.75. If the optical magnification indeed differs by $\\sim 30$\\% from the macrolens value, then the optically emitting region must be smaller, or at most a few times larger, than the Einstein ring scale. This accords nicely with (e.g.) the size of an accretion disk smaller than 100 Schwarzschild radii around a $10^9$M$_\\odot$ black hole (a scale of $3\\times 10^{16}$cm). This Einstein ring radius is only marginally, however, in accord with the apparent constancy of the microlensed magnification ratio: Since the Earth, the microlensing star (or stars, the effect being collective), and the quasar each have (3-dimensional) peculiar velocities of at least 300 km/s, the Earth should move through $\\sim 100$\\% microlensing variations in $\\sim 10$ yr (see Kochanek, Kolatt, and Bartelmann, 1996 for related calculations). So, the observed microlensing is about a factor 10 {\\it too constant}, and one is invited to speculate on whether something other than luck is the reason. Another invitation to speculation is the fact that Kundi\\'c et al. obtain rather different magnification ratios in their r- and g-band data, with the r ratio being $1.22\\pm 0.02$, with, again, the error bar depending on the method of analysis used. By any interpretation of the error bars, however, the r and g results are strongly discrepant. (Again note that there is no assumption that the fluctuations themselves have the same amplitude in the two colors, but only that the magnification {\\it ratio} should be the same.) Either a full $0.1$ mag of r-band galaxy light has escaped Kundi\\'c et al.'s careful subtraction in the B image, or something else is going on in the lens magnification ratio. With these two hovering peculiarities (possible excess time-constancy, and possible non-achromaticity), it seems useful to try to get additional information on the magnification ratio. This paper therefore asks the questions: Is the optical magnification ratio the same for the source region that produces the {\\it fluctuations} in quasar light as it is for the source region that produces the {\\it constant} light? And, does the magnification ratio of the fluctuations (which we may call the ``AC'' magnification ratio or ``ACMR'') agree more closely with the r- or g-band magnification ratio previously measured (here called the ``DC'' magnification ratio or ``DCMR'')? The answers to these questions can help diagnose the following situations: (i) If, as is true in many models, the size of the emitting region is much smaller than the microlensing scale, then all the magnification ratios should have the same value. (ii) If there is a problem with r-band galaxy subtraction -- or any other constant source of flux added to one lens component and not the other -- then the ACMR should represent the ``true'' microlens magnification ratio, and we might further expect it to be close to the g-band DCMR (where galaxy subtraction is a much smaller effect). (iii) If the optically emitting quasar accretion disk has a scale comparable to the microlensing scale, and has (as seems almost inevitable) color gradients, then the r and g ACMRs, and r and g DCMRs, might all be distinct. Indeed, the two ACMRs and two DCMRs then provide four distinct windows on the convolution of the accretion disk source with the microlensing pattern. Conceptually, one measures a DCMR and an ACMR as follows: Shift one of the light curves (A,B) in time by $\\tau$ to undo the lens delay. Fit each light curve by a constant value plus a residual time-varying part. The ratio of the constant values is the DCMR. Now, for the two time-varying residuals, fit for a model that makes the B residual a constant times the A residual. The best-fitting constant is the ACMR. This conceptual formulation, while simple, is actually not quite right. In the next section, we will give a statistical formulation of the problem that is more complete, and also more directly applicable to unevenly sampled data. In Section 3, we discuss some implementation details, and in Section 4 we apply the formulation to the published data of Kundi\\'c et al. (1995, 1997). Section 5 is discussion and conclusions. ", "conclusions": "While these data, in this analysis, do not support any very definitive conclusions, we may make the following remarks: Occam's razor would seem to indicate that the r-band light curve of Kundi\\'c et al. has about $0.1$ mag of residual, unsubtracted, constant light, as perhaps from unmodeled small-scale variations in the lens galaxy surface brightness. If this is the case, then all the data are compatible with a single magnification ratio for both colors and for both the fluctuating and constant pieces. This in turn suggests an accretion disk scale much smaller than the microlensing scale, in accord with theoretical prejudice. The utility of the ACMRs is that, taken together, they strongly favor the hypothesis that the g-band magnification ratio is the correct one, and that nothing more exotic is going wrong. We note, however (per E. Turner, private communication), that the galaxy G1 is something like 2 magnitudes fainter than component B in r band; thus the amount of unsubtracted light would need to be comparable to the total brightness of G1, which seems quite unlikely. It is up to the observers, not us, to decide whether Occam's razor should rule in this case. If $0.1$ mag of residual is not possibly present, then we must conclude that the accretion disk scale is comparable to the microlensing scale, and that the constant r-band part of the disk is {\\it more} strongly magnified than (at least some of) the other three regions. It seems likely on physical grounds that the fluctuating regions should be smaller than the constant regions, and that the g-band regions should be smaller than the r-band regions (temperature decreasing outward in the disk). For the larger (r-band and constant) region to have a higher magnification ratio than an enclosed smaller region, the larger region must extend to a place where the magnification is a superlinear function of position in the sky. This might indicate at least a fair chance of the B image passing through a caustic in the near ($\\sim 10$ year) future. This possibility, as well as the reconciliation of the relative constancy of the magnification ratio over the last 15 years, will be explored by Monte Carlo simulations in another paper (Press and Kochanek, in preparation)." }, "9803/astro-ph9803008_arXiv.txt": { "abstract": "We obtain restrictions on the universal baryon fraction, $f_{\\rm B} \\equiv \\Omega_{\\rm B}/\\Omega_0$, by assuming that the observed microlensing events towards the Large Magellanic Cloud are due to {\\em baryonic} MACHOs in the halo of the Galaxy and by extracting a bound to the total mass of the Milky Way from the motion of tracer galaxies in the Local Group. We find a lower bound $f_{\\rm B} > 0.29^{+0.18}_{-0.15}$. Consistency with the predictions of primordial nucleosynthesis leads to the further constraint on the total mass density, $\\Omega_0 \\alt 0.2$. ", "introduction": "It is a Herculean task to inventory the contents of the Universe (e.g., \\cite{fhp}). A more modest goal might be to pin down the baryonic fraction of the total mass, $f_{\\rm B}$ (e.g., \\cite{wnef}, \\cite{xrc}). If objects can be identified which are likely to provide a ``fair\" sample of $f_{\\rm B}$, we may avoid the daunting prospect of having to identify all the guises baryons may assume. Large clusters of galaxies offer a very promising site (\\cite{wnef}; \\cite{xrc}; \\cite{xray1}; \\cite{xray2}). To test the estimates of the systematic errors in $f_{\\rm B}$ derived from X-ray cluster data, it would be of value to measure $f_{\\rm B}$ in a completely different system, provided a case could be made that it will provide a ``fair\" sample. Suppose, for example, we could estimate the baryonic mass associated with the Galaxy. If we could also measure the corresponding ``dynamical\" mass, we could obtain an independent estimate of $f_{\\rm B}$ whose systematic uncertainties (and dependence on the Hubble parameter) differ from those which accompany the X-ray cluster determinations. In this paper we focus on the Local Group of galaxies (LG), using the MACHO mass estimates (\\cite{MACHO}) for a lower bound on the baryonic mass and relying on LG dynamics to constrain the total mass estimate. Microlensing experiments (\\cite{MACHO}) suggest that roughly half the mass in the halo of our Galaxy, out to the distance of the Large Magellanic Cloud (LMC), may be in the form of Massive Compact Halo Objects (MACHOs). One can imagine several exotic possibilities for the nature of the MACHOs. They could be very dense clusters of non-baryonic dark matter with special properties that allow them to clump inside their Einstein ring radii (\\cite{kt94}), or they could be primordial black holes. Neither of these possibilities is especially well motivated and each has its intrinsic difficulties, but neither can be excluded a priori. Stellar remnants such as old white dwarfs\\footnote{Neutron stars and black holes of stellar origin cannot constitute a significant halo fraction in view of the constraints arising from the observed metallicity and helium abundances (\\cite{ros90}).} appear to offer a more natural candidate (\\cite{MACHO}) which, however, is not without its problems too [e.g., white dwarfs require a rather narrow initial mass function in order to avoid overproducing low-mass stars or supernovae (\\cite{AL96})]. Dense and cold baryonic gas clouds have also been considered as a viable alternative for the observed gravitational microlenses (\\cite{cbc}; \\cite{cbc2}). Finally, it must be kept in mind that the observed microlensing may be due to objects which are not in the halo of the Galaxy. If the MACHOs are, indeed, stellar remnants (or cold baryonic gas clouds) in the halo of the Galaxy, then the mass of baryons within 50 kpc of the Galactic center is $M_{\\rm B} (50~{\\rm kpc}) \\geq M_{\\rm MACHO} = 2.0^{+1.2}_{-0.7} \\times 10^{11}M_\\odot$ (\\cite{MACHO}). The purpose of the present paper is to extract information on the universal baryon fraction from this number assuming the MACHOs are revealing baryonic matter in the Galaxy halo, and from the dynamics of the Local Group of galaxies. The constraint we obtain may be compared to the one derived from X-ray galaxy clusters (see, e.g., \\cite{xrc}; \\cite{xray1}; \\cite{xray2}), but it relies on different observations in a completely different physical system on a vastly different scale and, interestingly, has a different dependence on the Hubble parameter ($H_0 \\equiv 100h $~km~s$^{-1} $~Mpc$^{-1}$). The value of $M_{\\rm B} (50~{\\rm kpc})$ derived from microlensing experiments is approximately 50\\% of the total mass of the Galaxy out to this distance. The latter mass, presumably the sum of baryons and cold dark matter, is derived dynamically (see, e.g., \\cite{koch}). However, on the basis of this we cannot conclude that the primordial baryon fraction is $f_{\\rm B} \\approx 0.5$. Baryons are ``strongly'' interacting particles, while for the (non-baryonic) cold dark matter all interactions except gravitational can be neglected. Consequently, the density profile of the baryonic matter does not necessarily follow the density profile of the cold dark matter, and baryonic matter may be more (or less) concentrated towards the center of the gravitational well. However, we may be able to estimate the primordial baryon fraction if we take the ratio of baryons (as revealed by the MACHOs) to the total mass on some larger scale, which should be sufficiently large so that the matter inflow or outflow across the boundary of the region is negligible. The total mass of matter residing in such a larger region can be found dynamically; however, we cannot measure the mass of baryons separately on such larger scales. Although the baryonic halo may be expected to extend outside of the 50 kpc scale (in the form, e.g., of MACHOs, diffuse gas, satellite Galaxies, etc.), by neglecting these extended baryons we can obtain a lower bound on $f_{\\rm B}$. Indeed, while in the past there might have been violent processes of baryon ejection from the Galaxy accompanying, e.g., supernova explosions, analogous ejecta of cold dark matter is not expected. Therefore, by neglecting the unknown ejected component of baryons we will be on the ``safe side'' in our inequality for $f_{\\rm B}$, which, we emphasize, does rely on our assumption that MACHOs are baryonic matter in the halo of the Galaxy. For the larger reference scale we can choose the current turnaround radius for the LG. Initially, every shell of the Galaxy's building material expands with the Universe. Gradually, this expansion slows down and eventually a gravitationally bound shell separates from the general expansion. This shell stops expanding and then collapses (\\cite{gg72}). The radius of this first stopping point is the turnaround radius. With the passage of time shells that are more and more distant and less and less bound turn around sequentially, i.e., the turnaround radius propagates outward with time (for details see, e.g., \\cite{si}; \\cite{si2}; \\cite{stw}, 1997). There is one shell that is turning around now, at present; the corresponding distance of this shell from the center of mass of the system is the current turnaround radius. Collisionless cold (non-baryonic) dark matter is restricted to remain within this radius, which is just what we want for the larger reference scale. This picture of infall is valid independent of the assumption of spherical symmetry (the turnaround sphere will become a turnaround surface); for the model to be tractable analytically, we do assume spherical infall. ", "conclusions": "There remain several uncertainties in our LG baryon fraction estimate. One possibility which would weaken or even eliminate our constraint is if some of the observed microlensing events towards the LMC were due to an intervening satellite galaxy between us and the LMC, or due to debris in the LMC tidal tail (\\cite{z96}; \\cite{z97}). However, the MACHO collaboration concluded (\\cite{fg}) that if the lenses were in a foreground galaxy, it must be a particularly dark galaxy; see also (\\cite{AG97}). Moreover, the first observation of a microlensing event in the direction of the Small Magellanic Cloud (SMC) (\\cite{smc}), implies an optical depth in this direction roughly equal to that in the direction of the LMC. This makes it unlikely that a dwarf galaxy or a stellar stream between us and the LMC is responsible simultaneously for the observed microlensing towards the LMC and the SMC (\\cite{fg}; \\cite{AG97}). Recently, however, Gates et al. (1997) found Galactic models which explain the current microlensing data by a dark extension of the thick disk, reducing the MACHO fraction. It is to be anticipated that as more microlensing data are accumulated, these uncertainties will be resolved. We note that even in the absence of baryonic MACHOs there is still a limit, albeit much weaker, to $f_{\\rm B}$ from LG dynamics. The mass of baryons in the disk of the Galaxy provides a lower bound to $M_{\\rm B}$ which is smaller by a factor of $\\sim 3$ than the microlensing estimate we have used (\\cite{fhp}). Our lower bound to $f_{\\rm B}$ would be reduced by this factor while our upper bound to $\\Omega_0$ would be increased by the same factor. In summary, if the observed microlensing events are the result of baryonic MACHOs in the Galaxy halo, then the dynamics of the LG may be used to infer a {\\it lower} bound to the universal baryonic mass fraction: $f_{\\rm B} > 0.29^{+0.18}_{-0.15}\\, t_{10}^2$. If primordial nucleosynthesis is used to provide an {\\it upper} bound to the present baryonic density, we obtain an {\\it upper} bound to the present total mass density: $\\Omega_0 \\alt 0.2$ (with an extereme upper bound derived using nucleon-to-photon ratio based on the lithium abundance being $\\Omega_0 \\alt 0.47$)." }, "9803/astro-ph9803187_arXiv.txt": { "abstract": "We write a non-relativistic Lagrangian for a hierarchical universe. The equations of motion are solved numerically and the evolution of the fractal dimension is obtained for different initial conditions. We show that our model is homogeneous at the time of the last scattering, but evolves into a self-similar universe with a remarkably constant fractal dimension. We also show that the Hubble law is implied by this model and make an estimate for the age of the universe. \\vfill\\eject ", "introduction": "\\indent This is a decisive time for cosmology since our theories for the large scale structure of the universe are being seriously challenged by the ever-growing amount of data. The CfA1 redshift survey (de Lapparent, Geller \\& Huchra, 1986, 1988) was the first to reveal structures such as filaments and voids on scales where a random distribution of matter was expected. The most remarkable feature of these structures is the so-called ``great wall\" which is a coherent sheet of galaxies extended over an area of at least $60\\times 170 $ Mpc (Geller \\& Huchra 1989). Later on, deep pencil beam surveys (Broadhurst et al. 1990), the redshift surveys based on IRAS catalogue (Efstathiou et al. (1990a, 1990b), Saunders et al. 1991, Fisher et al 1996), the deep wide angle survey SSRS (de Costa 1988, 1994) and some others have shown inhomogeneities at scales where the galaxy-galaxy and cluster-cluster correlations were believed to be negligible. Of particular significance for the future are the two extensive redshift surveys, SLOAN and 2dF, about to commence, which aim to trace the 3-dimensional distribution of over one million galaxies across the northern and southern skies. The first quantitative study of the cosmic inhomogeneity lead to the well-known $1.8$ power law behaviour of the galaxy-galaxy correlation function (Groth \\& Peebles 1977, Peebles 1980, Davis \\& Peebles (1983a, 1983b)). Although this law has been consistently identified in different catalogues, the break away from it at larger scales and a crossover to homogeneity has not yet been established (Davis 1996, Pietronero 1987, Coleman \\& Pietronero 1992, Pietronero 1996). Whether there is a crossover to homogeneity or not, the power law nature of the two-point and higher-order correlation functions is itself suggestive of some kind of scaling behaviour at least in some range. The simplest structure that obeys such a scaling law is a single fractal. A theoretical model describing a non-analytic inhomogeneous scale-invariant universe is non-existent. The most-extensively-studied inhomogeneous cosmological model is Tolman spacetime (Tolman 1934, Bondi 1947). Tolman's dust solution has been used to model a hierarchical cosmology compatible with the observational analysis of the redshift surveys (Bonnor 1974, Ribeiro (1992a, 1992b), 1993). Recently, the Einstein equation for a scale-invariant spherically symmetric inhomogeneous, but isotropic, universe, which allows a non-vanishing pressure has been solved (Abdalla \\& Mohayaee 1997). However, the results obtained in this way are perturbative, assume a preferred center for the universe and violate the linearity of the Hubble law. A self-similar universe avoiding such difficulties can only be constructed for a non-analytic distribution of matter. It is rather a difficult task to construct a fractal metric and to solve Einstein equation for a self-similar universe. However, many cosmological phenomena can be accurately described by the Newtonian gravity, especially in the present matter-dominated era. In this work, we construct a self-similar universe whose dynamics is governed by the Newtonian gravity. We divide the universe into $k$ spherical clusters each of which contains $k$ subclusters which in their turn contain $k$ sub-subclusters. This clustering cascades down all the way to the level of the galaxies which are at the lowest rung on the clustering ladder. The mass and radius of each cluster can be used to define the fractal dimension of our model. We write the kinetic and potential energies of each cluster in terms of its center of mass energy and the internal energies of its subclusters. The thermal energy is obtained by requiring the total entropy of the canonical ensemble of the clusters and their subclusters to remain constant. The final Lagrangian is formulated in terms of two dynamical parameters: radius of the largest cluster and its ratio to the radius of its subclusters. The radius of the largest and smallest clusters, the ratio of the mass to the critical mass contained in a sphere of radius $20$ Mpc, the number of subclusters in each cluster and the ratio that characterises the relative significance of thermal and gravitational energies are left as free parameters. By fixing these to different observational values, we are able to solve the equations of motion numerically using a Pascal program. From the solutions, we can trace the evolution of the fractal dimension and verify the linearity of the velocity-distance relationship over time scales comparable to the age of the universe. The results are remarkable. We observe that for different initial conditions a nearly homogeneous universe with a fractal dimension close to 3 evolves into a universe with a fractal dimension of the order of 2 at the present time. This fractal dimension fluctuates slightly about the value of 2 over the future times but remains on the average constant. We also show that for insignificant thermal energies, the Hubble law is closely obeyed by our model. We also make an estimate for the age of our self-similar universe. This is one of the most challenging problems of Friedmann cosmology since the observed age of the old stars in the globular clusters is far bigger than the value estimated for the age of the universe in the nearly flat standard model. The age of the universe obtained in our model is related to the radius at which the crossover to homogeneity occurs. The farther the crossover radius the older is the universe. This article is organized as follows. In Section 2, we formulate our clustering model. In Section 3, we obtain the kinetic, the potential and the thermal energies, write the Lagrangian and the equations of motion. In Section 4, we solve these equations numerically and discuss the validity and limitations of our Newtonian approximation. In Section 5, we show different plots of the fractal dimension for different choices of the initial condition. Hubble law is discussed in Section 6. In Section 7, we study the evolution of the scale factor and obtain a value for the age of the universe in our model. Section 8 is devoted to the conclusion. ", "conclusions": "\\indent We have constructed a model for a nonrelativistic fractal universe. Our model starts off homogeneous and evolves rapidly to a self-similar universe with a remarkably constant fractal dimension of about 2. The homogeneity at the earlier times explains the isotropy of the microwave background radiation. We have also shown that the Hubble law is closely obeyed by our model for small thermal energies. We have estimated the age of the universe and have shown that it complies with the corresponding observational results. It remains an open problem to extend our model to multi-fractals and to the relativistic regime. {\\bf Acknowledgements} We thank R. Mansouri and M. Khorrami for useful discussions. This work has been partially supported by Conselho Nacional de Desenvolvimento Cient\\'\\i fico e Tecnol\\'ogico, CNPq, Brazil, and Funda\\c c\\~ao de Amparo \\`a Pesquisa do Estado de S\\~ao Paulo (FAPESP), S\\~ao Paulo, Brazil." }, "9803/astro-ph9803134_arXiv.txt": { "abstract": "We have completely mapped the Galactic globular cluster NGC~1851 with large-field, ground-based $VI$ CCD photometry and pre-repair $HST$/WFPC1 data for the central region. The photometric data set has allowed a $V $ vs. $ (V-I) $ colour--magnitude diagram for $\\sim$ 20500 stars to be constructed. From the apparent luminosity of the horizontal branch (HB) we derive a true distance modulus $(m-M)_0$ = 15.44 $\\pm$ 0.20. An accurate inspection of the cluster's bright and blue objects confirms the presence of seven ``supra-HB'' stars, six of which are identified as evolved descendants from HB progenitors. The HB morphology is found to be clearly bimodal, showing both a red clump and a blue tail, which are not compatible with standard evolutionary models. Synthetic Hertzsprung--Russell (HR) diagrams demonstrate that the problem could be solved by assuming a bimodal efficiency of the mass loss along the red giant branch (RGB). With the aid of Kolmogorov--Smirnov statistics we find evidence that the radial distribution of the blue HB stars is different from that of the red HB and subgiant branch (SGB) stars. We give the first measurement of the mean absolute $I$ magnitude for 22 known RR~Lyr variables ($ = 0.12 \\pm 0.20$ mag at a metallicity [Fe/H]~=~--1.28). The mean absolute $V$ magnitude is $ = 0.58 \\pm 0.20$ mag, and we confirm that these stars are brighter than those of the zero-age HB (ZAHB). Moreover, we found seven new RR~Lyr candidates (six $ab$ type and one $c$ type). With these additional variables the ratio of the two types is now $N_c$/$N_{ab} = 0.38$. From a sample of 25 globular clusters a new calibration for $\\Delta V_{\\rm bump}^{\\rm HB}$ as a function of cluster metallicity is derived. NGC~1851 follows this general trend fairly well. From a comparison with the theoretical models, we also find some evidence for an age--metallicity relation among globular clusters. We identify 13 blue straggler stars, which do not show any sign of variability. The blue stragglers are less concentrated than the subgiant branch stars with similar magnitudes for $r>80$ arcsec. Finally, a radial dependence of the luminosity function, a sign of mass segregation, is found. Transforming the luminosity function into a mass function (MF) and correcting for mass segregation by means of multi-mass King--Michie models, we find a global MF exponent $x_0=0.2\\pm 0.3$. ", "introduction": "Galactic globular clusters (GGC) are dynamically evolved objects. In order to understand the interplay between the internal dynamical processes and the influence of the Galactic potential, we must study a sample of GGCs comprising objects whose concentration, position in the Galaxy, luminosity and metallicity cover the whole observed range. The mass function and the radial profile must be determined for each cluster, in order to carry out a detailed dynamical analysis. The introduction of large-size CCDs has made this kind of investigations possible. With these detectors it is also possible to obtain deep photometry for the nearest globulars, and therefore to probe their mass functions over large mass intervals, in order to reach those MS stars which are more sensitive to dynamical effects (e.g. Pryor et al. 1986). A rich sample of stars is also essential in order to reveal and study the shortest-lived (and hence poorly known) phases of the stellar evolution (Renzini \\& Buzzoni 1986). Furthermore, the interactions between the single stars affect their evolution (e.g. Djorgovski et al. 1991). To establish the reliability of the stellar evolutionary models, we must therefore ascertain to what extent a GC colour--magnitude diagram and luminosity function is changed by the interactions among its stars. For the above reasons, our group started a project aimed at studying a number of globular clusters covering a wide range of the relevant parameters. NGC~1851 ($\\alpha_{2000} = 5^{\\rm h} 14^{\\rm m} 6^{\\rm s}.30$; $\\delta_{2000} = -40^\\circ 2\\arcmin 50\\farcs00$) has been selected for its position and its concentration. Its galactocentric distance, which is about twice that of the Sun, and its distance of 7.1~kpc from the Galactic plane (Djorgovski 1993) make it a typical halo object. Its concentration $c = 2.24$ is one of the highest in the list of Trager et al. (1995). A recent measurement of the cluster's proper motion has confirmed that NGC~1851 has halo-type kinematics (Dinescu et al. 1996). According to these authors, the space velocities of the cluster are $U=256\\pm35$~km s$^{-1}$, $V=-195\\pm26$~km s$^{-1}$, $W=-122\\pm30$~km s$^{-1}$, $\\Pi=195\\pm37$~km s$^{-1}$ and $\\Theta=167\\pm37$~km s$^{-1}$. Past photometric studies of the cluster are given in Alcaino (1969, 1971, 1976), Stetson (1981, hereafter S81), Sagar et al. (1988), Alcaino et al. (1990) and Walker (1992a, hereafter W92). The most exhaustive analysis is that of W92. His main results are that: (1) the cluster core, although unresolved, appears to be blue; (2) the HB is bimodal, showing both a red clump and an extended blue tail; (3) there are no radial trends in the relative numbers of red horizontal branch (RHB), blue horizontal branch (BHB) and red giant branch (RGB) stars for 48\\arcsec~ $< r <$ 190\\arcsec~; (4) the RGB ``bump'' is at $V$ = 16.15 $\\pm$ 0.03 mag; (5) the population ratio $R = N({\\rm HB})/N({\\rm RGB})$ has a value 1.26 $\\pm$ 0.10, which corresponds to a helium abundance $Y$ = 0.23 $\\pm$ 0.01 (computed by means of the R-method; e.g. Renzini 1977); (6) there are six blue straggler (BS) stars and six supra-RHB stars $[$15.7 mag $<$ $V$ $<$ 16.0 mag; 0.6 $< (B-V) <$ 0.8$]$ in the region 120\\arcsec~ $< r <$ 220\\arcsec~, and there is evidence of segregation only for the BS stars, so an origin for supra-RHB stars from BS stars is not supported by W92 data; (7) no significant abundance spread is found from the colour width of the main sequence (MS); and finally (8) an age of 14 $\\pm$ 1 Gyr results from the $\\Delta$~($B$$-$$V$) method (Sarajedini \\& Demarque 1990; VandenBerg et al. 1990). We have now obtained new {\\sl large-field} CCD $V$, $I$ photometry for NGC~1851. The new data set makes it possible to re-analyse the stellar content of the cluster with a much richer sample and, for the first time, allows a comprehensive study of its dynamical properties. However, the central regions of the cluster cannot be studied with this ground-based material, due to the extreme crowding of the core. To overcome this limitation, pre-repair {\\it Hubble Space Telescope} ({\\it HST}) images have been retrieved from the archives and reduced in order to sample the central stellar content, in particular the radial distribution of the HB stars. For the sake of comparison, the photometric catalogue of W92 has been also used. \\begin{figure}[t] \\psfig{figure=n1851_neg_.ps,width=8.8cm} \\caption[]{ The observed NTT/EMMI fields, sketched over a POSS field.} \\label{field_map} \\end{figure} \\begin{small} \\begin{table}[t] \\caption[]{Log of NTT/EMMI observations.} \\label{observs} \\begin{tabular}{cccccccc} \\noalign{\\smallskip} \\hline \\hline \\noalign{\\smallskip} Nr. & Field & $t_{\\rm exp}$(s) & Filter & Date & FWHM $[\\arcsec]$ \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} 1 & 6 & 50 & $V$ & 1993 Feb 18 & 1.2 \\\\ 2 & 6 & 70 & $I$ & 1993 Feb 18 & 1.2 \\\\ &&&&&\\\\ 3 & 5 & 45 & $I$ & 1993 Feb 18 & 1.2 \\\\ 4 & 5 & 10 & $I$ & 1993 Feb 18 & 1.4 \\\\ 5 & 5 & 10 & $V$ & 1993 Feb 18 & 1.1 \\\\ 6 & 5 & 30 & $V$ & 1993 Feb 18 & 1.3 \\\\ &&&&&\\\\ 7 & 2 & 30 & $V$ & 1993 Feb 18 & 1.0 \\\\ 8 & 2 & 40 & $I$ & 1993 Feb 18 & 1.2 \\\\ &&&&&\\\\ 9 & 3 & 60 & $I$ & 1993 Feb 18 & 1.2 \\\\ 10 & 3 & 44 & $V$ & 1993 Feb 18 & 1.1 \\\\ &&&&&\\\\ 11 & 1 & 45 & $V$ & 1993 Feb 18 & 1.0 \\\\ 12 & 1 & 60 & $I$ & 1993 Feb 18 & 1.2 \\\\ &&&&&\\\\ 13 & 4 & 60 & $I$ & 1993 Feb 18 & 1.1 \\\\ 14 & 4 & 45 & $V$ & 1993 Feb 18 & 1.1 \\\\ &&&&&\\\\ 15 & 7 & 45 & $V $ & 1993 Feb 18 & 1.1 \\\\ 16 & 7 & 60 & $I$ & 1993 Feb 18 & 1.2 \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} 17 & 8 & 70 & $I$ & 1993 Feb 19 & 1.1 \\\\ 18 & 8 & 55 & $V$ & 1993 Feb 19 & 1.1 \\\\ &&&&&\\\\ 19 & 9 & 55 & $V$ & 1993 Feb 19 & 1.2 \\\\ 20 & 9 & 70 & $I$ & 1993 Feb 19 & 1.2 \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} 21 & back & 120 & $V$ & 1993 Dec 10 & 1.0 \\\\ 22 & back & 180 & $I$ & 1993 Dec 10 & 1.1 \\\\ \\noalign{\\smallskip} \\hline \\end{tabular} \\end{table} \\end{small} ", "conclusions": "\\label{thediscussion} We have presented new large-field CCD photometry for $\\sim20500$ stars in the Galactic halo globular cluster NGC~1851, from both groundbased observations and a pre-repair {\\it HST} field. The photometric catalogue has been used to build a $V$ vs. ($V$--$I$) CMD, which has been analysed in detail. An extensive comparison of our data set with the predictions of the stellar models has also been performed. The effects of the dynamical evolution over the main sequence mass function have been investigated by means of a completeness-corrected luminosity function and the radial-count profile. \\paragraph{The evolved stellar content} With an accurate inspection of the cluster bright-blue objects, and a comparison with the numbers predicted from the background field and the Galactic count models, we have confirmed the presence of seven ``supra-HB'' stars in the CMD of NGC 1851. We have shown that six of the ``supra-HB'' stars could be evolved descendants from HB progenitors (post-HB or planetary nebulae). We have shown that standard evolutionary models are not able to reproduce the observed bimodal distribution of stars along the HB. Synthetic HR diagrams demonstrate that the problem could be solved by assuming that the efficiency of the RGB mass loss actually encompasses values going from 0.25 to 0.48. We have found evidence that the radial distribution of the blue HB stars is different from that of red HB and SGB stars. The BHB stars are significantly more concentrated than the SGB stars for $r>100$ arcsec. Though this distribution cannot be easily interpreted in terms of dynamical evolution, it might be related to the anomalous distribution of the BSs (see below). All the 27 known variable stars have been identified, and 26 have been measured in both colours (the remaining one being saturated). Twenty-two of them are RR~Lyr variables. For the first time, our photometry has allowed the mean absolute $I$ magnitude of the RR~Lyr variables to be obtained at a metallicity [Fe/H]~=~--1.28. The RR~Lyr are brighter than the ZAHB in the $V$ band, in accordance with the relation given by Carney et al. (1992). The positions and the photometry for seven new RR~Lyr candidates have been given. With these additional variables the ratio of the two types is now $N_c$/$N_{ab} = 0.38$, which reduces the current estimate N$_c$/$N_{ab} = 0.47$ (Wehlau et al. 1982). Thirteen BS stars have been identified outside the inner 80 arcsec. They do not show any sign of variability. We have investigated the radial distribution of the BSS. For $r>80$ arcsec, the BSs are less concentrated than the SGB stars with the same $V$ magnitude. We argue that the distribution of the BSs in the outer envelope of NGC 1851 might be similar to the distribution found by Ferraro et al. (1997) for the BSs in the envelope of M3. We have considered a sample of 25 globular clusters and have derived a new calibration for the $\\Delta V_{\\rm bump}^{\\rm HB}$ parameter as a function of cluster metallicity, and we have found that NGC~1851 follows this general trend fairly well. From a comparison with the corresponding slopes predicted by the isochrones library from Bertelli et al. (1994), we have found that perhaps an age--metallicity relation actually exists among globular clusters, with the metal poorest possibly being older. \\paragraph{Dynamical status of NGC~1851} We have been able to derive a complete LF down to $V \\simeq 23.5$ mag for stars in the region 190\\arcsec~ $< r <$ 650\\arcsec~, and down to $V \\simeq 22$ mag in the region 120\\arcsec~ $< r <$ 189\\arcsec~. The external LF is steeper than the internal one, and we have interpreted this result as a sign of mass segregation. By using the most updated mass--luminosity relations we have obtained MFs which can be well fitted by power laws with distinct exponents $x$. The observed value for the external MF is $x = 1.52 \\pm 0.18$, which is steeper than the value $0.89 \\pm 0.20$ found for the internal one. The global MF has been determined correcting the two observed mass functions for the effects of mass segregation, as predicted by the multi-mass King--Michie model which best fits the observed light profile of NGC 1851. The two values for the slope of the MF are compatible with the model if a global MF exponent $x_0=0.2\\pm 0.3$ is adopted. This value for the global MF slope is marginally smaller (MF flatter) than what would be expected from the relation between the slope of the MFs and the position in the Galaxy and the metallicity of the cluster proposed by Djorgovski et al. (1993). This might indicate that NGC 1851 has had a stronger gravitational interaction with the Galactic disc than the average of the Galactic GCs with similar position and metallicity. \\paragraph{} The above results indicate that NGC~1851 is a cluster where the dynamical evolution has affected both its evolved and unevolved stellar content. While the single findings are not of high statistical significance (mostly due to the small size of the stellar samples), taken together they give a coherent picture. Stellar encounters have led to mass segregation, as shown by the MF, which is steeper and steeper going from external to internal regions. They have probably contributed to the creation of the observed group of blue straggler stars, and possibly have triggered the formation of a blue tail in the HB. The internal dynamics of NGC~1851 has therefore influenced the evolution of its stars, introducing effects not reproducible by standard models. In turn, the dynamical evolution induced by the external gravitational field of the Galaxy has also very probably contributed to the modification of the present-day stellar population of NGC 1851, as strongly suggested by the anomalously flat global mass function." }, "9803/astro-ph9803302_arXiv.txt": { "abstract": "The shape of the radial velocity and light curves of 24 long-period ($30 \\leq P \\leq 134$ d) Cepheids in the Magellanic Clouds shows a progression with the period. The sequences of the radial velocity and light curves are based only on a small sample of stars; however, evident changes of the shape can be seen in Cepheids with period between 90 and 134 d. The Fourier parameter--period diagrams for the radial velocity curves show trends which remind in part those of Cepheids with period near 10 d. The plausible interpretation is a resonance, probably $P_0/P_1=2$ between the fundamental and the first overtone mode. The possible importance of this phenomenon for the study of stellar structure and evolution in relatively far galaxies is emphasized \\footnote {Tables 2 and 3 are only available in electronic form at the CDS via anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5)}. ", "introduction": "\\begin{table*} \\caption[]{List of the analyzed long-period Cepheids in Magellanic Clouds} \\begin{flushleft} \\begin{tabular}{lllllllllll} \\hline\\noalign{\\smallskip} HV & $P_L$ & $N_L$ & Ref$_L$ & ord$_L$ & $\\sigma_L$ (mag) & $P_{RV}$ & $N_{RV}$ & Ref$_{RV}$ & ord$_{RV}$ & $\\sigma_{RV}$ (km~s$^{-1}$) \\\\ 904 L & 30.400 & 58 & 10 & 6 & .034 & & & & & \\\\ 1002 L & 30.4694 & 60 & 1 - 6 & 7 & .032 & & & & & \\\\ 899 L & & & & & & 31.027 & 49 & 7 & 7 & 1.7 \\\\ 2294 L & 36.530 & 98 & 2 - 6 & 8 & .032 & & & & & \\\\ 879 L & 36.817 & 31 & 4 & 5 & .048 & 36.782 & 43 & 7 & 6 & 1.4 \\\\ 909 L & 37.5828 & 61 & 1 - 4 & 4 & .047 & 37.510 & 64 & 7 & 5 & 1.1 \\\\ 11182 S & 39.1941 & 44 & 2,3,5 & 3 & .039 & & & & & \\\\ 2257 L & & & & & & 39.294 & 48 & 7 & 6 & 1.7 \\\\ 2195 S & 41.7988 & 39 & 1,2,5 & 5 & .049 & & & & & \\\\ 2338 L & 42.2153 & 55 & 2,4,5 & 7 & .048 & 42.184 & 59 & 7 & 8 & 1.4 \\\\ 837 S & 42.739 & 55 & 2 - 5 & 6 & .037 & 42.673 & 86 & 9 & 10 & 1.6 \\\\ 877 L & 45.107 & 78 & 2 - 5 & 3 & .041 & & & & & \\\\ 900 L & 47.5417 & 62 & 1 - 5 & 5 & .039 & 47.544 & 55 & 8 & 6 & 1.2 \\\\ 2369 L & 48.3311 & 108 & 1 - 6 & 7 & .034 & & & & & \\\\ 824 S & 65.8306 & 62 & 1,4 & 5 & .030 & 65.755 & 40 & 8 & 5 & 1.7 \\\\ 11157 S & & & & & & 69.06 & 62 & 9 & 5 & 1.2 \\\\ 834 S & 73.399 & 106 & 2 - 6 & 7 & .035 & 73.648 & 53 & 8 & 7 & 1.3 \\\\ 2827 L & 78.858 & 55 & 3,4 & 3 & .026 & 78.626 & 43 & 7 & 4 & 1.3 \\\\ 829 S & & & & & & 85.577 & 45 & 8 & 6 & 1.0 \\\\ 5497 L & 99.078 & 67 & 3 - 6 & 3 & .026 & 99.040 & 41 & 8 & 3 & 1.2 \\\\ 2883 L & 109.277 & 58 & 2 - 4,6 & 5 & .032 & 109.071 & 52 & 8 & 3 & 2.1 \\\\ 2447 L & 117.941 & 68 & 3 - 6 & 3 & .022 & 118.841 & 39 & 8 & 2 & 1.3 \\\\ 821 S & 127.490 & 117 & 1 - 6 & 4 & .040 & 127.083 & 50 & 8 & 4 & 1.4 \\\\ 883 L & 133.893 & 98 & 1 - 6 & 4 & .045 & 134.75 & 92 & 9 & 6 & 2.7 \\\\ \\noalign{\\smallskip} \\hline \\end{tabular} \\end{flushleft} Ref.: 1. Gascoigne \\& Kron (1965); 2. Madore (1975); 3. Van Genderen (1977, 1983); 4. Martin \\& Warren (1979); 5. Eggen (1977); 6. Freedman et al. (1985); 7. Imbert et al. (1985); 8. Imbert et al. (1989); 9. Imbert (1994); 10. Sebo \\& Wood (1995). \\end{table*} The longer period Cepheids are particularly important in the context of the primary distance scale because they are bright enough to be visibile at a great distance. However, due to their low number, they have been poorly studied both observationally and theoretically. Simon \\& Kanbur (\\cite{sk}) considered 50 Cepheids in Galaxy and IC 4182 with period $P$ less than 70 d, compared them with hydrodynamical pulsation models and concluded that a detailed comparison between theory and observations must await a more extensive and accurate sample of observed stars. Antonello \\& Morelli (\\cite{am}) analyzed all the available photometric $V$ data of galactic Cepheids with period less than 70 d looking for possible resonance effects; they noted some small features in the Fourier parameters--period diagrams which were ascribed tentatively to expected resonances. Aikawa \\& Antonello (\\cite{aa}) tried to reproduce these observations with nonlinear models, but their conclusion was that the increasing nonadiabaticity of the pulsation with period probably reduces the effectiveness of resonance mechanisms. Finally, Simon \\& Young (\\cite{sy}) studied long period Cepheids in the period range $10 \\leq P \\leq 50$ d in Magellanic Clouds looking for galaxy-to-galaxy differences in the Cepheid distributions. Resonances between pulsation modes, which were studied essentially in Cepheids with $P$ less than about 30 d, represent a powerful comparison tool between observations and theoretical model predictions, because they affect the shape of the curves of pulsating stars in specific period ranges. The comparison of the Fourier parameters of observed and theoretical light and radial velocity curves allows to probe the stellar interior and to put constraints on the stellar physical parameters. After the work of Simon \\& Lee (\\cite{sl}) on the resonance $P_0/P_2=2$ at $P_0 \\sim 10$ d between the fundamental and second overtone mode in classical bump Cepheids, several papers by various authors were devoted to this topic, from both the observational and theoretical point of view. For example, Buchler \\& Kovacs (\\cite{bk}) and Moskalik \\& Buchler (\\cite{mb}) studied the general effects of 2:1 and 3:1 resonances in radial stellar pulsations and discussed the possible astrophysical implications, Petersen (\\cite{pe}) discussed the possible two- and three-mode resonances in Cepheids, and Antonello (\\cite{an}) looked for the expected effects in short period Cepheids. Recent reviews on galactic and Magellanic Cloud Cepheids pulsating in fundamental and first overtone mode and on the problems raised by the comparison with the pulsational models are those by Buchler (\\cite{bu}) and Beaulieu \\& Sasselov (\\cite{bs}). The resonance effects in Magellanic Cloud Cepheids cannot be reproduced by models constructed using current input physics and reasonable mass--luminosity relations; in particular the case of first overtone Cepheids characterized by $P_1/P_4=2$ (Antonello et al. \\cite{apr}) has proven to be rather difficult for theorists. We mention in passing also the recent resonance $P_2/P_6=2$, studied in the models of hypothetical second overtone mode Cepheids (Antonello \\& Kanbur \\cite{ak}). In the present work we have considered the long-period Cepheids in Magellanic Clouds, with available photometric and radial velocity data which were suitable for Fourier decomposition. The initial purpose of the work was simply to extend the comparison between theory and observations to Cepheids with the longest known periods, but the probable discovery of a new resonance effect suggested to publish the present Letter in advance of the comparison with the hydrodynamical models (Antonello \\& Aikawa, in preparation). ", "conclusions": "The analysis of the radial velocity and light curves of the long-period Cepheids in the Magellanic Clouds indicates the presence of a progression which we interpret tentatively as an effect of the resonance $P_0/P_1=2$ between the fundamental and the first overtone mode. Since the longest period Cepheids are also the brightest, the present results could be of some importance for the study of stellar structure and evolution in far galaxies, because the stars with $P \\sim $ 100 d ($M_V \\sim -7$ mag) are about three magnitudes brighter than those at 10 d, in which the well known resonance $P_0/P_2=2$ is occurring. The disadvantage is the low number of such stars. Few Cepheids with $P > 80$ d have been found in relatively nearby galaxies (NGC 6822, IC 1613, NGC 300; see e.g. Madore \\cite{mm}), while in the Magellanic Clouds there are just a few of stars in comparison with a total of some thousand Cepheids. Presently the Hubble Space Telescope Key Project on the Extragalactic Distance Scale is optimized for the detection of Cepheids with period between 3 and 60 d (e.g. Ferrarese et al. \\cite{fe}), therefore it is not possible to derive reliable conclusions about the number of long-period stars. We just note that in NGC 925 Silbermann et al. (\\cite{sil}) found 4 stars with probable $P > 80$ d over a total of 80 Cepheids. Assuming that a sufficient number of such stars is detected and our interpretation is correct, the comparison of the observed resonance effect with nonlinear model predictions will allow to test the input physics and put constraints on the physical parameters of the stars in relatively far galaxies in the same way as it is occurring for the Galaxy and Magellanic Cloud Cepheids with shorter periods." }, "9803/astro-ph9803244_arXiv.txt": { "abstract": "The observed evolution of the galaxy cluster X-ray integral temperature distribution function between $z=0.05$ and $z=0.32$ is used in an attempt to constrain the value of the density parameter, $\\Omega_{0}$, for both open and spatially-flat universes. We estimate the overall uncertainty in the determination of both the observed and the predicted galaxy cluster X-ray integral temperature distribution functions at $z=0.32$ by carrying out Monte Carlo simulations, where we take into careful consideration all the most important sources of possible error. We include the effect of the formation epoch on the relation between virial mass and X-ray temperature, improving on the assumption that clusters form at the observed redshift which leads to an {\\em overestimate} of $\\Omega_0$. We conclude that at present both the observational data and the theoretical modelling carry sufficiently large associated uncertainties to prevent an unambiguous determination of $\\Omega_{0}$. We find that values of $\\Omega_{0}$ around 0.75 are most favoured, with $\\Omega_{0}<0.3$ excluded with at least 90 per cent confidence. In particular, the $\\Omega_{0}=1$ hypothesis is found to be still viable as far as this dataset is concerned. As a by-product, we also use the revised data on the abundance of galaxy clusters at $z=0.05$ to update the constraint on $\\sigma_8$ given by Viana \\& Liddle \\shortcite{VL}, finding slightly lower values than before. ", "introduction": "The number density of rich clusters of galaxies at the present epoch has been used to constrain the amplitude of mass density fluctuations on a scale of $8\\,h^{-1}\\,{\\rm Mpc}$ (Evrard 1989; Henry \\& Arnaud 1991; White, Efstathiou \\& Frenk 1993a; Viana \\& Liddle 1996, henceforth VL; Eke, Cole \\& Frenk 1996; Kitayama \\& Suto 1997). This is usually referred to as $\\sigma_{8}$, where $h$ is the present value of the Hubble parameter, $H_{0}$, in units of $100\\;{\\rm km}\\,{\\rm s}^{-1}\\,{\\rm Mpc}^{-1}$. However, the derived value of $\\sigma_{8}$ depends to a great extent on the present matter density in the Universe, parameterized by $\\Omega_{0}$, and more weakly on the presence of a cosmological constant, $\\Lambda$. The cleanest way of breaking this degeneracy is to include information on the change in the number density of rich galaxy clusters with redshift \\cite{FWED}, the use of X-ray clusters for this purpose having been proposed by Oukbir \\& Blanchard \\shortcite{OB} and subsequently further investigated \\cite{HM,OB97}. Several attempts have been made recently, with wildly differing results \\cite{Henry,FBC,Grossetal,BB,Ekeetal,Retal}. The best method to find clusters of galaxies is through their X-ray emission, which is much less prone to projection effects than optical identification. Further, the X-ray temperature of a galaxy cluster is at present the most reliable estimator of its virial mass. This can then be used to relate the cluster mass function at different redshifts, calculated for example within the Press--Schechter framework \\cite{PS,BCEK}, to the observed cluster X-ray temperature function. We can therefore compare the evolution in the number density of galaxy clusters seen in the data with the theoretical expectation for large-scale structure formation models, which depends significantly only on the assumed values of $\\Omega_{0}$ and $\\lambda_{0}\\equiv\\Lambda/3H^{2}_{0}$, the latter being the contribution of $\\Lambda$ to the total present energy density in the Universe. However, the X-ray temperature of a cluster of galaxies is not an easily measurable quantity, as compared to its X-ray luminosity. A minimum flux is required, so that there is a high enough number of photons for the statistical errors in the temperature determination to be reasonably small. Because of this, although estimates of the present-day cluster X-ray temperature function have been available since the early 90's \\cite{ESFA,HA}, the change in the cluster X-ray temperature function as we look further into the past has been much more difficult to determine. Estimates for the X-ray temperatures of individual clusters with redshifts as high as 0.3 have been available for some years (e.g.~see David et al.~1993), but only with the advent of the {\\em ASCA} satellite has it been possible to measure X-ray temperatures for clusters of galaxies around that redshift in a systematic way, and to go to even higher redshifts. The evolution of the cluster X-ray luminosity function with redshift, though easier to determine, provides much weaker constraints on $\\Omega_{0}$ and $\\lambda_{0}$, due to the fact that the X-ray luminosity of a galaxy cluster is not expected to be a reliable estimator of its virial mass (e.g.~Hanami 1993). Though it could in principle provide some indication of the change of the cluster X-ray temperature function with redshift, the problem is that not only is there considerable scatter in the present-day cluster X-ray temperature verses luminosity relation \\cite{Davetal,Fetal}, but it is also not known how the relation may change with redshift, though recently it has been argued that at least up to $z=0.4$ it does not seem to evolve \\cite{MScharf,AF,Retal,SBO}. The deepest complete X-ray sample of galaxy clusters presently available is the one obtained from the {\\em Einstein Medium Sensitivity Survey} ({\\em EMSS}) \\cite{Getal,Hetal}. This sample is restricted to objects with declination larger than $-40^{\\rm o}$ and is flux-limited, with $F_{{\\rm obs}}\\geq1.33\\times10^{-13}\\;{\\rm erg}\\,{\\rm cm}^{-2}\\,{\\rm s}^{-1}$, where $F_{{\\rm obs}}$ is the cluster flux in the 0.3 to 3.5 keV band which falls in a $2'.4\\times2'.4$ {\\em EMSS} detect cell. It presently contains 90 clusters of galaxies, after a few misidentifications were recently removed \\cite{GioiaL,Netal}. This is the only complete galaxy cluster catalogue beyond a redshift of 0.3, and as such unique in providing the means to distinguish between different possible values for $\\Omega_{0}$ and $\\lambda_{0}$. However, until the recent effort by Henry \\shortcite{Henry}, very few X-ray temperatures were known for those galaxy clusters in the {\\em EMSS} sample with redshifts exceeding 0.15 (see Sadat et al.~1998 for a recent compilation). Henry \\shortcite{Henry} used {\\em ASCA} to observe all galaxy clusters in the {\\em EMSS} cluster sample with $0.3\\leq z \\leq 0.4$ and $F_{{\\rm obs}}\\geq2.5\\times10^{-13}\\;{\\rm erg}\\,{\\rm cm}^{-2}\\,{\\rm s}^{-1}$. The resulting sub-sample of 10 clusters has a median redshift of 0.32, and the data obtained for each cluster, the X-ray flux, luminosity and temperature, can be found in Table 1 of Henry (1997). We will use this data together with the present-day (median redshift 0.05) cluster X-ray temperature function. We work within the extended Press--Schechter formalism proposed by Lacey \\& Cole (1993, 1994), which allows an estimation of the formation times of dark matter halos. We will assume the dark matter to be cold, and consider the cases of an open universe, where the cosmological constant is zero, and a spatially-flat universe, such that $\\lambda_{0}=1-\\Omega_{0}$. ", "conclusions": "{}From the above analysis, we conclude that {\\em at present} it is not possible to reliably exclude any interesting value for $\\Omega_{0}$ on the basis of X-ray cluster number density evolution alone, due to the limited statistical significance of the available observational data and to uncertainties in the theoretical modelling of cluster formation and evolution. However, we do find that values of $\\Omega_{0}$ below 0.3 are excluded at least at the 90 per cent confidence level. Values of $\\Omega_{0}$ between 0.7 to 0.8 are those most favoured, though not strongly. These results are basically independent of the presence or not of a cosmological constant. Our conclusions support those of Colafrancesco, Mazzotta \\& Vittorio \\shortcite{CMV}, who tried to estimate the uncertainty involved in the estimation of the cluster X-ray temperature distribution function at different redshifts based on its present-day value. They found this uncertainty, given the still relatively poor quality of the data, to be sufficiently large to preclude the imposition of reliable limits on the value of $\\Omega_{0}$. Our results disagree with those of Henry \\shortcite{Henry} and Eke et al.~\\shortcite{Ekeetal}, as they found the preferred $\\Omega_{0}$ to lie between 0.4 to 0.5, with the $\\Omega_{0}=1$ hypothesis strongly excluded. This disagreement is mainly the consequence of our focus on the threshold X-ray temperature of 6.2 keV, while they draw their conclusions based on the analysis of the results obtained for several threshold X-ray temperatures. Further below we will repeat our calculations assuming a threshold X-ray temperature of 4.8 keV, and we will find that when we calculate the joint probability of some value for $\\Omega_{0}$ being excluded on the basis of the results concerning either one or both threshold X-ray temperatures of 6.2 keV and 4.8 keV, the favoured value for $\\Omega_{0}$ decreases to around 0.55. Some of the reasons for our choice of deriving the conclusions solely based on the results obtained for the 6.2 keV threshold were mentioned at the end of subsection 3.1 and others will be detailed below. Other less important contributions to the difference between our results and those presented by Henry \\shortcite{Henry} and Eke et al.~\\shortcite{Ekeetal} are the different assumed normalization for the virial mass to X-ray temperature relation, and the corrections in the expected values in the Universe for both $N(>6.2\\,{\\rm keV},\\,0.05)$ and $N(>6.2\\,{\\rm keV},\\,0.32)$ due to the uncertainties in the X-ray cluster temperature measurements. Note that changing the mean of the bootstrap distribution obtained for $N(>6.2\\,{\\rm keV},\\,0.32)$ to its theoretically-expected overall value in some $\\Omega_{0}$ universe and then calculating the exclusion level on the estimated value for $N(>6.2\\,{\\rm keV},\\,0.32)$ in the Universe given the dataset in Henry \\shortcite{Henry}, rather than just using the original bootstrap distribution to impose an exclusion level on the theoretically-expected overall value for $N(>6.2\\,{\\rm keV},\\,0.32)$ in that $\\Omega_{0}$ universe, does not seem to make much difference. This is a reflection of the fact that the bootstrap distributions recovered do not have a strongly asymmetric shape. Our disagreement with Eke et al.~\\shortcite{Ekeetal} on the level of exclusion of the $\\Omega_{0}=1$ hypothesis is also due to our much larger assumed uncertainty in the theoretically-expected overall value for $N(>6.2\\,{\\rm keV},\\,0.32)$. For the $\\Omega_{0}=1$ hypothesis to be favoured, one requires the lowest possible observed value for $N(>6.2\\,{\\rm keV},0.32)$. This is best achieved if, for the sample of 10 galaxy clusters used in its calculation, the X-ray temperatures turn out to be on average lower than the assumed mean, and the X-ray fluxes higher. A higher ratio between the extended and detect cell fluxes for the {\\em EMSS} at $z=0.32$ would also help. On the theoretical side, the higher one decides the expected value for $N(>6.2\\,{\\rm keV},0.32)$ is, the more compatible with the data the $\\Omega_{0}=1$ hypothesis becomes. This can be best achieved if, in decreasing order of importance, the cluster virial mass at fixed X-ray temperature is being underestimated, $\\delta_{{\\rm c}}$ is lower than the canonical value 1.7 and $f$, the assembled fraction of a cluster virial mass after which the X-ray temperature does not change significantly, is assumed greater than 0.75. However, the single most important factor in determining the theoretically-expected overall value for $N(>6.2\\,{\\rm keV},0.32)$ is the present-day normalization for the dispersion of the density field, $\\sigma_{8}$, which in turn results from the observational value for the present density $N(>6.2\\,{\\rm keV},\\,0.05)$. Although we worked with all X-ray clusters that make up the dataset in Henry \\shortcite{Henry}, and even estimated the effect of also considering the 5 clusters with lower X-ray fluxes present in the {\\em EMSS} in the redshift bin from 0.3 to 0.4, in fact we only used the abundance of clusters with X-ray temperatures in excess of 6.2 keV to constrain $\\Omega_{0}$. We mentioned some of the reasons for this choice in Section~3. Nevertheless, we decided to repeat the same calculations for a threshold X-ray temperature of 4.8 keV. This value also well represents the mean curve going through the observed cumulative X-ray temperature distribution function at both $z=0.05$ and $z=0.32$. \\begin{figure} \\centering \\leavevmode\\epsfysize=5.4cm \\epsfbox{zclus_fig4.eps}\\\\ \\caption[Figure4]{The absolute exclusion levels for different values of $\\Omega_{0}$ in both the open and spatially-flat cases, when the threshold X-ray temperature of 4.8 keV is used.} \\end{figure} The results regarding the best-fit value for $\\Omega_{0}$, presented in Figure~4, are somewhat different from those we obtained when the threshold X-ray temperature was assumed to be 6.2 keV. This is particularly true if the correction for the possibility of any of the 5 clusters with the lowest X-ray fluxes in the $0.30.8$ is excluded without the correction, being this limit lowered to 0.7 when the correction is included. One can also estimate the joint probability of some $\\Omega_{0}$ value being excluded on the basis of the results relative to either one or both X-ray temperature thresholds. Assuming the data used in the calculations for the two thresholds is independent, the results then imply that the favoured value for $\\Omega_{0}$ is close to 0.55 (0.50 if the incompleteness correction is included) and the $\\Omega_{0}=1$ hypothesis is excluded at the 99 per cent level. This agrees very well with the results of Henry \\shortcite{Henry} and Eke et al.~\\shortcite{Ekeetal}, leading us to believe that the main difference between our analysis and theirs is our decision to draw our conclusions solely based on the exclusion levels obtained for the X-ray temperature threshold of 6.2 keV. A further potential problem one must consider when working with clusters whose observed X-ray temperature is as low as 4.8 keV is the possibility that the energy in the intracluster gas has increased as a result of (pre-)heating by supernovae and starbursts in the cluster galaxies. In fact this is the leading hypothesis (e.g. Navarro, Frenk \\& White 1995; Markevitch 1998) put forward to explain the discrepancy between the observed slope of the X-ray temperature--luminosity relation, close to 0.3, and the expected value of 0.5 if clusters evolve in a self-similar way \\cite{Kaiser}. Following Eke et al.~\\shortcite{Ekeetal}, we assume that in a cluster whose observed X-ray temperature is 4.8 keV, 17 per cent of its energy, that is 0.8 keV per intracluster gas particle, was due to (pre-)heating produced by processes occurring inside the cluster galaxies. This is approximately the amount of energy that gets injected into the intracluster gas particles in the simulation of Metzler \\& Evrard \\shortcite{ME}, where a galaxy cluster's X-ray temperature, which would otherwise be 5.6 keV, increased to 6.4 keV. Note however that in the scheme proposed by Eke et al.~\\shortcite{Ekeetal} a cluster this large would not be (pre-)heated to the extent simulated by Metzler \\& Evrard \\shortcite{ME}, as in their proposal Eke et al.~\\shortcite{Ekeetal} assume that the energy gained by each intracluster gas particle due to (pre-)heating decreases as a galaxy cluster becomes larger, being close to zero for galaxy clusters with X-ray temperatures exceeding 6.2 keV. \\begin{figure} \\centering \\leavevmode\\epsfysize=5.4cm \\epsfbox{zclus_fig5.eps}\\\\ \\caption[Figure5]{The absolute exclusion levels for different values of $\\Omega_{0}$ only for the open case, when the threshold X-ray temperature of 4.8 keV is used. The full curve includes a correction (FC) for the possibility of the 5 clusters with lowest fluxes in the {\\em EMSS} located between $z=0.3$ and $z=0.4$ having X-ray temperatures in excess of 4.8 keV. The dashed curve includes a correction (HC) for the possibility of (pre-)heating of the intracluster medium due to processes within the cluster galaxies. The dotted curve includes both corrections.} \\end{figure} The above assumption means that the observed values for $N(>4.8\\,{\\rm keV},\\,z)$, when $z=0.05$ and $z=0.32$, should now be compared with the theoretically-expected values for $N(>4.0\\,{\\rm keV},\\,z)$ at those redshifts. The resulting exclusion levels on the value of $\\Omega_{0}$ can be seen in Figure 5 for the open case. There is little difference compared to the results in Figure 4 that follow from the standard no-heating calculation. The lower value for $\\sigma_{8}$, required to match theory and observations at $z=0.05$, more than compensates for the expected increase in the number of galaxy clusters with X-ray temperatures in excess of 4.8 keV at $z=0.05$, in effect bringing this number down. In fact, the standard no-heating calculation for a threshold X-ray temperature of 4.8 keV requires a value for $\\sigma_{8}(\\Omega_{0})$, so that the observed value for $N(>4.8\\,{\\rm keV},\\,0.05)$ is reproduced, that is less than 3 per cent below that required by the $>6.2$ keV data, quoted in equation (\\ref{final1}). On the other hand, including the (pre-)heating correction, the required $\\sigma_{8}(\\Omega_{0})$ value drops to 19 per cent below that preferred by the $>6.2$ keV data. Though the coincidence between the $\\sigma_{8}$ values obtained for the two X-ray temperature thresholds 4.8 keV and 6.2 keV under the no-heating assumption may be accidental, it could indicate that (pre-)heating was relatively unimportant at least for the galaxy clusters observed at $z=0.05$ with X-ray temperatures exceeding 4 keV. If (pre-)heating was more important in the past than today, then the required $\\sigma_{8}(\\Omega_{0})$ value would be that obtained through the standard no-heating hypothesis, but the comparison at $z=0.32$ would include the (pre-)heating correction. This would push the theoretically-expected value for $N(>4.8\\,{\\rm keV},\\,0.32)$ up, favouring higher values for $\\Omega_{0}$. This is not as far-fetched as it may seem, given that it is well known that the star-formation rate peaks before $z=1$ (e.g. Madau, Ferguson \\& Dickinson 1998; Baugh et al. 1998), and consequently so does the rate of supernovae Type II (the rate of supernovae Type Ia peaks a few Gyr later) and the probability of starbursts. The results for the 4.8 keV threshold X-ray temperature are close to those found by Eke et al.~\\shortcite{Ekeetal}, leading us to believe that their exclusion levels for $\\Omega_{0}$ are dominated by the information associated with the threshold X-ray temperatures 4.0 keV and 5.0 keV. In our view, the analysis for these X-ray temperature thresholds carries with it a sufficient number of uncertainties, due to the problems mentioned above, so as to render the constraints imposed on $\\Omega_{0}$ not very trustworthy. Only the data regarding clusters with X-ray temperatures in excess of about 6 keV seems sufficiently free of modelling problems so as to be potentially useful in constraining $\\Omega_{0}$. Another possible complication has arisen from recent work by Blanchard, Bartlett and Sadat \\shortcite{BBS} who use a sample of 50 galaxy clusters with mean redshift of 0.05, which were identified through the {\\em ROSAT} satellite, to estimate the cumulative X-ray temperature distribution function at $z=0.05$. They claim the number density of galaxy clusters at $z=0.05$ with X-ray temperatures exceeding 4 keV is being {\\em underestimated} when the Henry \\& Arnaud cluster sample is used. Through the X-ray cumulative temperature distribution function at $z=0.05$ they obtain, they then estimate $\\Omega_{0}$ using the {\\em EMSS} cluster abundance in the redshift bin $0.30.4$ the best fit to the data is given by a model of clustering evolution with a comoving \\rn = 2.37 \\Mpc\\ and $\\epsilon = -0.4^{+0.37}_{-0.65}$, consistent with published measures of the clustering evolution. To match the canonical value of \\rn = 5.4 \\Mpc, found for the clustering of local galaxies, requires a value of $\\epsilon = 2.10^{+0.43}_{-0.64}$ (significantly more than linear evolution). The log likelihood of this latter fit is 4.15 less than that for the \\rn = 2.37 \\Mpc\\ model. We, therefore, conclude that the parameterization of the clustering evolution of $(1+z)^{-(3+\\epsilon)}$ is not a particularly good fit to the data. ", "introduction": "The evolution of the clustering of galaxies as a function of redshift provides a sensitive probe of the underlying cosmology and theories of structure formation. In an ideal world we would measure the spatial correlation function of galaxies as a function of redshift and type and use this to compare with the predictions of different galaxy formation theories. Observationally, however, our ability to efficiently measure galaxy spectra falls rapidly as a function of limiting magnitude and consequently we are limited to deriving spatial statistics from small galaxy samples and at relatively bright magnitude limits (e.g.\\ $I_{AB} < 22.5$, Le Fevre et al.\\ 1996, Carlberg et al.\\ 1997). To increase the size of the galaxy samples and thereby reduce the shot noise the standard approach has been to measure the angular correlation function, i.e.\\ the projected spatial correlation function (Brainerd et al.\\ 1996, Woods and Fahlman 1997). While this allows us to extend the measure of the clustering of galaxies to fainter magnitude limits ($R<29$, Villumsen et al.\\ 1997) it has an associated limitation. For a given magnitude limit the amplitude of the angular correlation function is sensitive to the width of the galaxy redshift distribution, N(z). At faint magnitude limits N(z) is very broad and consequently the clustering signal is diluted due to the large number of randomly projected pairs. In this letter we introduce a new approach for quantifying the evolution of the angular correlation function; we apply photometric redshifts (Connolly et al.\\ 1995, Lanzetta et al.\\ 1996, Gwyn and Hartwick 1996, Sawicki et al.\\ 1997) to isolate particular redshift intervals. In so doing we can remove much of the foreground and background contamination of galaxies and measure an amplified angular clustering. We discuss here the particular application of this technique to the Hubble Deep Field (HDF; Williams et al.\\ 1996). ", "conclusions": "Photometric redshifts provide a simple statistical means of directly measuring the evolution of the clustering of galaxies. By isolating narrow intervals in redshift space we can reduce the number of randomly projected pairs and detect the clustering signal to high redshift and faint magnitude limits. Applying these techniques to the HDF we can characterize the evolution of the angular 2 pt correlation function out to $z=1.6$. For redshifts $0.4>R_p$ are disfavored. In such models, the observed spectrum would have an exponential cutoff, unless a source is accidentally close to the observer. In the latter case the flux would be strongly anisotropic. Finally, in many cases TD give UHECR fluxes lower than the observed ones. We showed here that this is the case for monopole-string networks. Superconducting and ordinary cosmic strings probably belong to this category as well, although some loopholes still remain to be closed. With all these constraints taken into account, it appears that only necklaces, monopolonium and relic SH particles survive as potential UHE sources. The most important observational signature of TD as sources of UHE CR is the presence of photon-induced EAS. For all known mechanisms of UHE particle production the pions (and thus photons) dominate over nucleons. At energies lower than $1\\cdot 10^{12}~GeV$, protons have considerably larger attenuation length than photons and the observed proton flux can be dominant. Nevertheless, even in this case photons reach an observer from sources located inside the sphere of radius $R_\\gamma (E)$ (assuming that $R_\\gamma >D$). Unlike protons, photons propagate rectilinearly, indicating the direction to the sources. {\\em Necklaces} with a large value of $r=m/\\mu d > 10^7$ have a small separation $D 1\\cdot 10^{10}~GeV$, as compared with the direction perpendicular to the Galactic Plane. A flux from the Virgo cluster might be another signature of this model. The search for photon induced showers is not an easy experimental task. It is known (see e.g. Ref.\\cite{AK}) that in the UHE photon-induced showers the muon content is very similar to that in proton-induced showers. However, some difference in the muon content between these two cases is expected and may be used to distinguish between them observationally. A detailed analysis would be needed to determine this difference. The Landau-Pomeranchuk-Migdal (LPM) effect \\cite{LPM} and the absorption of photons in the geomagnetic field are two other important phenomena which affect the detection of UHE photons \\cite{AK,Kasa}; (see \\cite{ps} for a recent discussion). The LPM effect reduces the cross-sections of electromagnetic interactions at very high energies. However, if the primary photon approaches the Earth in a direction characterized by a large perpendicular component of the geomagnetic field, the photon is likely to decay into electron and positron \\cite{AK,Kasa}. Each of them emits a synchrotron photon, and as a result a bunch of photons strikes the Earth atmosphere. The LPM effect, which strongly depends on energy, is thus suppressed. If on the other hand a photon moves along the magnetic field, it does not decay, and LPM effect makes shower development in the atmosphere very slow. At extremely high energies the maximum of the showers can be so close to the Earth surface that it becomes \"unobservable\" \\cite{ps}. We suggest that for all energies above the GZK cutoff the showers be analyzed as candidates for being induced by UHE photons, with the probability of photon splitting in the geomagnetic field determined form the observed direction of propagation, and with the LPM effect taken into account. The search for photon-induced showers can be especially effective in the case of Fly's Eye detector which can measure the longitudinal development of EAS. The future Auger detector will have, probably, the highest potentiality to resolve this problem." }, "9803/astro-ph9803101_arXiv.txt": { "abstract": "We searched for cluster X-ray luminosity and radius evolution using our sample of 201 galaxy clusters detected in the 160~deg$^2$ survey with the \\ROSAT\\/ PSPC (Vikhlinin et al.\\ 1998). With such a large area survey, it is possible, for the first time with \\ROSAT\\/, to test the evolution of luminous clusters, $L_x>3\\times10^{44}\\,$\\ergpersec\\ in the 0.5--2~keV band. We detect a factor of 3--4 deficit of such luminous clusters at $z>0.3$ compared to the present. The evolution is much weaker or absent at modestly lower luminosities, 1--$3\\times10^{44}\\,$\\ergpersec. At still lower luminosities, we find no evolution from the analysis of the $\\log N - \\log S$ relation. The results in the two upper $L_x$ bins are in agreement with the {\\em Einstein\\/} EMSS evolution result (Gioia et al.\\ 1990a, Henry et al.\\ 1992) while being obtained using a completely independent cluster sample. The low-$L_x$ results are in agreement with other \\ROSAT\\/ surveys (e.g.\\ Rosati et al.\\ 1998, Jones et al.\\ 1998). We also compare the distribution of core radii of nearby and distant ($z>0.4$) luminous (with equivalent temperatures 4--7~keV) clusters, and detect no evolution. The ratio of average core radius for $z\\sim0.5$ and $z<0.1$ clusters is $0.9\\pm0.1$, and the core radius distributions are remarkably similar. A decrease of cluster sizes incompatible with our data is predicted by self-similar evolution models for high-$\\Omega$ universe. ", "introduction": "The cluster evolution rate is a strong test of cosmological parameters (e.g., White \\& Rees 1978, Kaiser 1986, Eke, Cole \\& Frenk 1996). It is best to study evolution using X-ray selected samples of distant clusters which are much less affected by projection than the optically selected samples (van Haarlem et al.\\ 1997). Of all the interesting cluster parameters such as mass, velocity dispersion, and temperature, the X-ray luminosity is the most accessible to measurements with present-day instruments, and most of the earlier studies were focused on evolution of the cluster X-ray luminosity function. A strong evolution of cluster luminosities at $z\\sim0.1$ was reported from the EXOSAT survey (Edge et al.\\ 1990), but was later disproved by the \\ROSAT\\/ All-Sky Survey (Ebeling et al.\\ 1997). At higher redshifts, negative evolution of the cluster X-ray luminosity function was first reported by Gioia et al.\\ (1990a) using the {\\em Einstein}\\/ Extended Medium Sensitivity Survey (EMSS; Gioia et al.\\ 1990b, Stocke et al.\\ 1991). Gioia et al.\\ and later Henry et al.\\ (1992) compared the cluster luminosity functions below and above $z=0.3$. They found that while the number of the low luminosity clusters does not evolve, there is a significant deficit of luminous, $L_x(0.3-3.5\\mbox{~keV})>5\\times10^{44}\\,$\\ergpersec, clusters at high redshift. This EMSS result was questioned recently. Nichol et al.\\ (1997) reanalyzed the EMSS cluster sample using \\ROSAT\\/ X-ray and new optical observations and argued that the evolution reported in the original EMSS papers was not significant. Several groups pursued independent searches for distant clusters in archival \\ROSAT\\/ PSPC observations. Collins et al.\\ (1997) found that the redshift distribution of 35 clusters detected in their 17~deg$^2$ survey is consistent with no evolution. This contradicted the earlier claim by Castander et al.\\ (1995) of a strong evolution in a similar sample; however, the latter authors used an X-ray source detection algorithm not optimized for the cluster search. Jones et al.\\ (1998) presented the $\\log N - \\log S$ relation for 46 clusters from their 16~deg$^2$ survey and found that this relation is consistent with no evolution of the $L_x<2\\times10^{44}\\,$\\ergpersec\\ (0.5--2~keV band) clusters. Rosati et al.\\ (1998) derived cluster luminosity functions up to $z\\sim 0.8$ from their sample of 70 clusters detected in a 33~deg$^2$ survey, and found no evolution at low luminosities, $L_x<3\\times10^{44}\\,$\\ergpersec. However, none of these \\ROSAT\\/ surveys covers an area large enough to probe the evolution of the luminous clusters, and their no-evolution claims do not contradict the EMSS results. Our 160~deg$^2$ survey (Vikhlinin et al.\\ 1998, hereafter Paper~I) is the first \\ROSAT\\/ survey comparable with the EMSS in sky coverage for distant clusters. We are able to test, and confirm, the EMSS evolution results even with the incomplete redshift data currently at hand. Eventually, when the spectroscopic work is complete, we will be able to characterize the luminosity evolution more accurately. In this Letter, we also show that the cluster X-ray core radii do not evolve between $z\\sim0.5$ and now. Throughout the paper, we use definitions $f_{-14}$ and $L_{44}$ for flux and luminosity in the 0.5--2~keV energy band in units of $10^{-14}\\,$\\ergs\\ and $10^{44}\\,$\\ergpersec, respectively. We also use $H_0=50$~km~s$^{-1}$~Mpc$^{-1}$ and $q_0=0.5$. ", "conclusions": "We present a first \\ROSAT\\/ analysis of the evolution of luminous, $L_x>3\\times10^{44}\\,$\\ergpersec\\ distant clusters. We find a significant, factor of 3--4, decrease in the number of such clusters at $z>0.3$, confirming the detection of evolution in the EMSS (Gioia et al.\\ 1990a, Henry et al.\\ 1992). At lower luminosities, 1--$3\\times10^{44}\\,$\\ergpersec, the evolution is undetectable, with a decrease in number by a factor of only $1.3\\pm0.2$. This is also consistent with the EMSS and other \\ROSAT\\/ surveys (e.g.\\ Rosati et al.\\ 1998). The absence of evolution of low luminosity clusters is also supported by the analysis of the $\\log N - \\log S$ distribution from which we find that the cluster volume emissivity, dominated by low-luminosity objects, does not evolve. The observed evolution can be reproduced by a model in which the characteristic luminosity decreases with redshift, but the comoving number density of clusters increases. Such models arise naturally in the hierarchical cluster formation scenario (e.g.\\ Kaiser 1986). We compare the distribution of core-radii of distant, $z>0.4$ and nearby clusters. We find that the distribution of core radii at $z>0.4$ is very similar to that in nearby clusters; the average radius has changed at $z>0.4$ by a factor of only $0.9\\pm0.1$. A stronger change is expected for hierarchical cluster formation in a flat universe (Kaiser 1986, Cen \\& Ostriker 1994). We also note that the assumption of no evolution of cluster sizes has been essentially used in flux measurements and area calculations in several X-ray surveys (e.g.\\ EMSS, Nichol et al. 1997), but is only verified here for the first time. \\bigskip \\centerline{\\includegraphics[width=3.25in]{axdistr.ps}} \\vskip -15pt \\figcaption{Core-radius distribution for distant, $z>0.4$, clusters, derived from our survey (solid and dashed histogram for $q_0=0.5$ and $q_0=0$, respectively). The shaded histogram shows the core-radius distribution for nearby luminous clusters from Jones \\& Forman (1998). The angular resolution limit of our survey (15\\arcsec) corresponds to 120~kpc at the median redshift of distant clusters, well below the peak of the distribution. \\label{fig:axdistr}} \\medskip" }, "9803/astro-ph9803337_arXiv.txt": { "abstract": "If the theoretical relationship between white dwarf mass and orbital period for wide-orbit binary radio pulsars is assumed to be correct, then the neutron star mass of PSR J2019+2425 is shown to be $\\sim 1.20 M_{\\odot}$. Hence the mass of the neutron star in this system prior to the mass transfer phase is expected to have been $< 1.1 M_{\\odot}$. Alternatively this system descends from the accretion induced collapse (AIC) of a massive white dwarf.\\\\ We estimate the magnetic inclination angles of all the observed wide-orbit low-mass binary pulsars in the Galactic disk using the core-mass period relation and assuming that the spin axis of an accreting neutron star aligns with the orbital angular momentum vector in the recycling process of the pulsar. The large estimated magnetic inclination angle of PSR J2019+2425, in combination with its old age, gives for this system evidence against alignment of the magnetic field axis with the rotational spin axis. However, in the majority of the similar systems the distribution of magnetic inclination angles is concentrated toward low values (if the core-mass period relation is correct) and suggests that alignment has taken place. ", "introduction": " ", "conclusions": "" }, "9803/astro-ph9803294_arXiv.txt": { "abstract": "s{ Samples of high-redshift galaxies are easy to select in the millimetre/submillimetre (mm/submm) waveband using sensitive telescopes, because their flux density--redshift relations are expected to be flat, and so the selection function is almost redshift-independent at redshifts greater than 0.5. Source counts are expected to be very steep in the mm/submm waveband, and so the magnification bias due to gravitational lensing is expected to be very large, both for lensing by field galaxies and for lensing by clusters. Recent submm-wave observations of lensed images in clusters have constrained the submm-wave counts directly for the first time. In the next ten years our knowledge of galaxy evolution in this waveband will be greatly enhanced by the commissioning of sensitive new instruments and telescopes, including the CMBR imaging space mission {\\em Planck Surveyor}. This paper highlights the important features of gravitational lensing in the submm waveband and discusses the excellent prospects for lens searches using these forthcoming facilities. } ", "introduction": " ", "conclusions": "\\begin{enumerate} \\item The surface density of distant galaxies in the submm waveband, and therefore the expected surface density of gravitational lenses and the effects of source confusion in future observations, is now known with reasonable accuracy. \\item The {\\em Planck Surveyor} survey and surveys using other forthcoming mm/submm-wave telescopes will produce catalogues of distant sources that will be of great interest for studies of galaxy evolution. Lensed galaxies and quasars will be detected with an efficiency of up to 10\\% in these surveys, and a sample of order 1000 lenses could be compiled. \\end{enumerate}" }, "9803/hep-ph9803418_arXiv.txt": { "abstract": "The neutrino capture rate measured by the Russian-American Gallium Experiment is well below that predicted by solar models. To check the response of this experiment to low-energy neutrinos, a 517 kCi source of \\nuc{51}{Cr} was produced by irradiating 512.7~g of 92.4\\%-enriched \\nuc{50}{Cr} in a high-flux fast neutron reactor. This source, which mainly emits monoenergetic 747-keV neutrinos, was placed at the center of a 13.1 tonne target of liquid gallium and the cross section for the production of \\nuc{71}{Ge} by the inverse beta decay reaction $\\mnuc{71}{Ga}(\\nu_e,e^-)\\mnuc{71}{Ge}$ was measured to be {[5.55 \\+/-~0.60~(stat) \\+/-~0.32~(syst)]} \\E{-45} cm$^2$. The ratio of this cross section to the theoretical cross section of Bahcall for this reaction is 0.95 \\+/- 0.12 (expt) $^{+0.035}_{-0.027}$ (theor) and to the cross section of Haxton is 0.87 \\+/- 0.11 (expt) \\+/- 0.09 (theor). This good agreement between prediction and observation implies that the overall experimental efficiency is correctly determined and provides considerable evidence for the reliability of the solar neutrino measurement. ", "introduction": "Gallium experiments are uniquely able to measure the principal component of the solar neutrino spectrum. This is because the low threshold of 233 keV \\cite{Audi and Wapstra 95} for inverse beta decay on the 40\\% abundant isotope \\nuc{71}{Ga} is well below the end point energy of the neutrinos from proton-proton fusion, which are predicted by standard solar models to be about 90\\% of the total flux. The Russian-American Gallium Experiment (SAGE) has been measuring the capture rate of solar neutrinos with a target of gallium metal in the liquid state since January 1990. The measured capture rate \\cite{Abdurashitov et al. 94,Gavrin 98} is 67~\\+/-7~(stat)~$^{+5}_{-6}$ (syst) SNU\\footnote{1 SNU corresponds to one neutrino capture per second in a target that contains 10$^{36}$ atoms of the neutrino absorbing isotope.}, a value that is well below solar model predictions of 137 $^{+8}_{-7}$ SNU \\cite{Bahcall and Pinsonneault and Wasserburg 95} and 125~\\+/- 5 SNU \\cite{Turck-Chieze and Lopes 93}. In addition, the GALLEX Collaboration, which has been measuring the solar neutrino capture rate with an aqueous GaCl$_3$ target since 1991, observes a rate of 70~\\+/- 7~$^{+4}_{-5}$ SNU \\cite{Hampel et al. 96}. The other two operating solar neutrino experiments, the chlorine experiment \\cite{Cleveland et al. 98} and the Kamiokande experiment \\cite{Suzuki et al. 95}, have significantly higher-energy thresholds, and thus are not able to see the neutrinos from $pp$ fusion. When the results of these four solar neutrino experiments are considered together, a contradiction arises which cannot be accommodated by current solar models, but which can be explained if one assumes that neutrinos can transform from one species to another \\cite{Bahcall 94,Berezinsky 94,Parke 95,Hata et al. 94,Castellani et al. 94,Bahcall et al. 95,Heeger and Robertson 96}. The gallium experiment, in common with other radiochemical solar neutrino experiments, relies on the ability to extract, purify, and count, all with well known efficiencies, a few atoms of a radioactive element that were produced by neutrino interactions inside many tonnes of the target material. In the case of 60 tonnes of Ga, this represents the removal of a few tens of atoms of \\nuc{71}{Ge} from $5 \\times 10^{29}$ atoms of Ga. To measure the efficiency of extraction, about 700 $\\mu$g of stable Ge carrier is added to the Ga at the beginning of each exposure, but even after this addition, the separation factor of Ge from Ga is still 1 atom in 10$^{11}$. This impressively stringent requirement raises legitimate questions about how well the many efficiencies that are factored into the final result are known. It has been understood since the outset that a rigorous check of the entire operation of the detector (i.e., the chemical extraction efficiency, the counting efficiency, and the analysis technique) would be made if it is exposed to a known flux of low-energy neutrinos. In addition to verifying the operation of the detector, such a test also eliminates any significant concerns regarding the possibility that atoms of \\nuc{71}{Ge} produced by inverse beta decay may be chemically bound to the gallium (so-called ``hot atom chemistry'') in a manner that yields a different extraction efficiency than that of the natural Ge carrier. In other words, it tests a fundamental assumption in radiochemical experiments that the extraction efficiency of atoms produced by neutrino interactions is the same as that of carrier atoms. This article describes such a test, in which a portion of the SAGE gallium target was exposed to a known flux of \\nuc{51}{Cr} neutrinos and the production rate of \\nuc{71}{Ge} was measured. Similar tests have also been made by GALLEX \\cite{Hampel et al. 98}. Although a direct test with a well-characterized neutrino source lends significant credibility to the radiochemical technique, we note that numerous investigations have been undertaken during the SAGE experiment to ensure that the various efficiencies are as quoted \\cite{Abdurashitov et al. 94}. The extraction efficiency has been determined by a variety of chemical and volumetric measurements that rely on the introduction and subsequent extraction of a known amount of the stable Ge carrier. A test was also carried out in which Ge carrier doped with a known number of \\nuc{71}{Ge} atoms was added to 7 tonnes of Ga. Three standard extractions were performed, and it was demonstrated that the extraction efficiencies of the carrier and \\nuc{71}{Ge} follow each other very closely. Another experiment was performed to specifically test the possibility that atomic excitations might tie up \\nuc{71}{Ge} in a chemical form from which it would not be efficiently extracted. There is a concern that this might occur in liquid gallium because the metastable Ga$_2$ molecule exists with a binding energy of \\about1.6 eV. In this experiment the radioactive isotopes \\nuc{70}{Ga} and \\nuc{72}{Ga} were produced in liquid gallium by neutron irradiation. These isotopes quickly beta decay to \\nuc{70}{Ge} and \\nuc{72}{Ge}. The Ge isotopes were extracted from the Ga using our standard procedure and their number was measured by mass spectrometry. The results were consistent with the number expected to be produced based on the known neutron flux and capture cross section, thus suggesting that chemical traps are not present. This experiment is not conclusive, however, because the maximal energy imparted to the \\nuc{70}{Ge} nucleus following beta decay of \\nuc{70}{Ga} is 32 eV, somewhat higher than the maximal energy of 20 eV received by the \\nuc{71}{Ge} nucleus following capture of a 747-keV neutrino from \\nuc{51}{Cr} decay (and considerably higher than the maximum nuclear recoil energy of 6.1 eV after capture of a 420-keV neutrino from proton-proton fusion). Further evidence that the extraction efficiency was well understood came from monitoring the initial removal from the Ga of cosmogenically produced \\nuc{68}{Ge}. This nuclide was generated in the Ga as it resided on the surface exposed to cosmic rays. When the Ga was brought underground, the reduction in the \\nuc{68}{Ge} content in the initial extractions was the same as for the Ge carrier. These numerous checks and auxiliary measurements have been a source of confidence in our methodology, yet it is clear that a test with an artificial neutrino source of known activity provides the most compelling validation of radiochemical procedures. This article is an elaboration of work that previously appeared in Ref.~\\cite{Abdurashitov et al. 96}. The experimental changes since the previous Letter are some minor refinements in the selection of candidate \\nuc{71}{Ge} events and in the treatment of systematic errors; recent cross section calculations are also included. The central experimental result given here is almost identical to what was reported earlier. ", "conclusions": "The primary motivation for the \\nuc{51}{Cr} source experiment was to determine if there is any unexpected problem in either the chemistry of extraction or the counting of \\nuc{71}{Ge}, i.e., to see if there is some unknown systematic error in one or both of the efficiency factors in $\\epsilon$, the product of extraction and counting efficiencies. If some such systematic error were to exist, then the value of $\\epsilon$ that we have used in the preceding will be in error by the factor $E$, defined as $E \\equiv \\epsilon_{\\text{true}}/ \\epsilon_{\\text{measured}}$. Since the cross section is inversely proportional to $\\epsilon$, this hypothetical error is equivalent to the cross section ratio, $E = \\sigma_{\\text{measured}}/\\sigma_{\\text{true}}$. An experimental value for $E$ can be set from our measured cross section, Eq.~(\\ref{cross section result}), if one assumes that the true cross section is equivalent to the theoretically calculated cross section. Then $E \\approx R \\equiv \\sigma_{\\text{measured}}/\\sigma_{\\text{theoretical}}$. Neutrino capture cross sections averaged over the four neutrino lines of \\nuc{51}{Cr} have been calculated by Bahcall \\cite{Bahcall 97} and by Haxton \\cite{Haxton 98}. Bahcall, assuming that the strength of the two excited states in \\nuc{71}{Ge} that can be reached by \\nuc{51}{Cr} neutrinos is accurately determined by forward-angle $(p,n)$ scattering, gives a result of 5.81 (1.0 $^{+0.036}_{-0.028})$ \\E{-45} cm$^2$. The upper limit for the uncertainty was set by assuming that the excited state strength could be in error by as much as a factor of 2; minor contributions to the uncertainty arise from forbidden corrections, the \\nuc{71}{Ge} lifetime, and the threshold energy. An independent consideration of the contribution of excited states has been made by Hata and Haxton \\cite{Hata and Haxton 95} and very recently by Haxton \\cite{Haxton 98}. They argue that, because of destructive interference between weak spin and strong spin-tensor amplitudes in \\nuc{71}{Ge}, the strengths determined from $(p,n)$ reactions are, for some nuclear levels, poor guides to the true weak interaction strength. In particular, Haxton finds the weak interaction strength of the $(5/2)^-$ level in \\nuc{71}{Ge} at an excitation energy of 175 keV to be much greater than the value that is measured by the $(p,n)$ scattering reaction, and calculates a total \\nuc{51}{Cr} cross section of (6.39~\\+/- 0.68) \\E{-45} cm$^2$. This cross section was deduced from the measured $(p,n)$ cross sections for the two excited states, and uses a large-basis shell model calculation to correct for the presence of spin-tensor contributions. Since not all known theoretical uncertainties were included, the stated error here is a lower bound. Combining our statistical and systematic uncertainties for the cross section in quadrature into an experimental uncertainty, we can thus give estimates for $E$: \\begin{eqnarray} E & \\approx & R \\equiv \\frac{\\sigma_{\\text{measured}}} {\\sigma_{\\text{theoretical}}} \\\\ & = & \\left\\{ \\begin{array}{l l} 0.95 \\pm 0.12 \\text{ (expt)}\\ ^{+0.035}_{-0.027} \\text{ (theor)} & \\text{ (Bahcall)}, \\\\ 0.87 \\pm 0.11 \\text{ (expt)}\\ \\pm 0.09 \\text{ (theor)} & \\text{ (Haxton)}. \\end{array} \\right. \\nonumber \\end{eqnarray} \\noindent With either of these theoretical cross sections, $R$ is consistent with unity, which implies that the total efficiency of the SAGE experiment to the neutrinos from \\nuc{51}{Cr} is close to 100\\%. The measurement reported here should not be interpreted as a direct calibration of the SAGE detector for solar neutrinos. This is because the \\nuc{51}{Cr} neutrino spectrum differs from the solar spectrum, there is a 10\\%--15\\% uncertainty in the theoretical value for the \\nuc{51}{Cr} cross section, and the total experimental efficiency for each solar neutrino measurement is known to a higher precision than the 12\\% experimental uncertainty obtained with the \\nuc{51}{Cr} source. As a result, the solar neutrino measurements reported by SAGE should not be scaled by the factor $E$. Rather, we consider the Cr experiment as a test of the experimental procedures, and conclude that it has demonstrated {\\it with neutrinos} that there is no unknown systematic uncertainty at the 10\\%--15\\% level. The neutrino spectrum from \\nuc{51}{Cr} is very similar to that of \\nuc{7}{Be}, but at slightly lower energy. Since the response of \\nuc{71}{Ga} to \\nuc{7}{Be} neutrinos is governed by the same transitions that are involved in the \\nuc{51}{Cr} source experiment, we can definitely claim that, if the interaction strength derived from the \\nuc{51}{Cr} experiment is used in the analysis of the solar neutrino results, then the capture rate measured by SAGE includes the full contribution of neutrinos from \\nuc{7}{Be}. This observation holds independent of the value of $E$ or of cross section uncertainties. This demonstration is of considerable importance because a large suppression of the \\nuc{7}{Be} neutrino flux from the sun is one consequence of the combined analysis of the four operating solar neutrino experiments \\cite{Bludman et al. 93,Akhmedov et al. 95}. GALLEX has completed two \\nuc{51}{Cr} measurements whose combined result, using the cross section of Bahcall \\cite{Bahcall 97}, can be expressed as $R$ = 0.93~\\+/- 0.08 \\cite{Hampel et al. 98}, where the uncertainty in the theoretical cross section has been neglected. Both SAGE and GALLEX, which employ very different chemistries, give similar results for the solar neutrino capture rate and have tested their efficiencies with neutrino source experiments. The solar neutrino capture rate measured in Ga is in striking disagreement with standard solar model predictions and there is considerable evidence that this disagreement is not an experimental artifact." }, "9803/astro-ph9803011_arXiv.txt": { "abstract": "We describe narrowband and spectroscopic searches for emission-line star forming galaxies in the redshift range 3 -- 6 with the 10 m Keck\\,II Telescope. These searches yield a substantial population of objects with only a single strong (equivalent width $\\gg 100$ \\AA) emission line, lying in the $4000 - 8500$ \\AA\\ range. Spectra of the objects found in narrowband--selected samples at $\\lambda \\sim5390$ \\AA\\ and $\\sim6741$ \\AA\\ show that these very high equivalent width emission lines are generally redshifted Ly$\\alpha\\ \\lambda\\,1216$ \\AA\\ at $z\\sim3.4$ and 4.5. The density of these emitters above the 5$\\sigma$ detection limit of $1.5\\times 10^{-17}$ ergs cm$^{-2}$ s$^{-1}$ is roughly 15,000/$\\sq^{\\circ}$/unit $z$ at both $z\\sim3.4$ and 4.5. A complementary deeper ($1\\ \\sigma \\sim 10^{-18}$ ergs cm$^{-2}$ s$^{-1}$) slit spectroscopic search covering a wide redshift range but a more limited spatial area ($200\\ \\sq''$) shows such objects can be found over the redshift range $z=3 - 6$, with the currently highest redshift detected being at $z=5.64$. The Ly$\\alpha$ flux distribution can be used to estimate a minimum star formation rate in the absence of reddening of roughly $0.01\\ M_{\\odot}$ Mpc$^{-3}$ yr$^{-1}$ ($H_0 = 65\\ {\\rm km}\\ {\\rm s}^{-1}\\ {\\rm Mpc}^{-1}$, $q_0 = 0.5$). Corrections for reddening are likely to be no larger than a factor of two, since observed equivalent widths are close to the maximum values obtainable from ionization by a massive star population. Within the still significant uncertainties, the star formation rate from the Ly$\\alpha$--selected sample is comparable to that of the color-break--selected samples at $z\\sim 3$, but may represent an increasing fraction of the total rates at higher redshifts. This higher-$z$ population can be readily studied with large ground-based telescopes. ", "introduction": "The search for high-redshift galaxies and the effort to map the star formation history of galaxies have progressed rapidly in the last several years as magnitude--limited spectroscopic surveys pushed into the $z=1-3$ range (Cowie et al.\\markcite{large_sample} 1996; Cohen et al.\\markcite{cohenhdf} 1996), while color--based selection techniques produced many objects in the $z=2.5-5$ range (Steidel et al.\\markcite{stei96a} 1996a, \\markcite{stei96b}1996b; Franx et al.\\markcite{franx} 1997; Steidel et al.\\markcite{stei98} 1998), particularly in the exquisite Hubble Deep Field (HDF) sample (Lowenthal et al.\\markcite{low97} 1997). However, the galaxies chosen by these techniques correspond to objects with ongoing massive star formation and small amounts of extinction, and may represent only part of the populations at these early epochs. More evolved objects may be heavily dust--reddened and more easily picked out at submillimeter wavelengths (Smail, Ivison, \\& Blaine\\markcite{smail97} 1997; Hauser et al.\\markcite{dirbe} 1998), while earlier stages in evolution may have relatively little continuum light and be too faint to be seen in the magnitude--limited samples or selected with the color-break techniques at current sensitivity limits. This latter class of objects may represent the earliest stages of the galaxy formation process, in which substantial amounts of metals have yet to form. These galaxies may have much stronger Ly$\\alpha$ emission relative to the stellar continuum, since they have massive star formation that can excite the Ly$\\alpha$ emission line, but without so much dust that the line is suppressed, and this can result in very high observed equivalent widths in the range of 100--200\\,(1+$z$) \\AA\\ (e.g., Charlot \\& Fall\\markcite{charl93} 1993). Such objects may be hard to pick out with color-break techniques but be detectable in Ly$\\alpha$ searches of sufficient depth (Cowie\\markcite{cowie88} 1988; Thommes\\markcite{thommes} 1996). An increased incidence of strong Ly$\\alpha$ emission does, indeed, appear in the color-break samples at fainter continuum magnitudes (Steidel et al.\\markcite{stei98} 1998). Earlier blank-field Ly$\\alpha$ surveys (e.g., Thompson, Djorgovski, \\& Trauger\\markcite{tdt95} 1995; Thompson \\& Djorgovski\\markcite{td95} 1995) failed to find such objects, as Pritchet\\markcite{pri94} (1994) has summarized, but their sensitivity lay at the margin of where such objects would be expected in significant numbers (Cowie\\markcite{cowie88} 1988; Thommes \\& Meisenheimer\\markcite{thommes95} 1995). However, Hu \\& McMahon\\markcite{br2237} (1996), using very deep targeted narrowband searches ($1\\ \\sigma = 1.5 \\times 10^{-17}$ ergs cm$^{-2}$ s$^{-1}$), found $z\\sim4.55$ Ly$\\alpha$-emitting galaxies with the very strong emission and weak or undetected continuua predicted for early star-forming objects. The advent of 10 m telescopes, along with this successful detection of Ly$\\alpha$ emitters, prompted us to undertake a new survey that has been successful in detecting blank-field high-$z$ Ly$\\alpha$ emitters. The present Letter describes the early results of this search, which uses extremely deep narrowband filter exposures taken with LRIS (Oke et al.\\markcite{lris} 1995) on the Keck II telescope to search for emission-line populations at extremely faint flux levels ($1\\ \\sigma=3 \\times 10^{-18}$ ergs cm$^{-2}$ s$^{-1}$). This survey picks out sources of extremely high equivalent width ($W_{\\lambda}>100$ \\AA) emission lines as candidates, and then uses followup LRIS spectroscopy (\\S2) to determine if these are Ly$\\alpha$ emitters. The first results from the Hawaii survey with a 5390/77 \\AA\\ filter, corresponding to Ly$\\alpha$ emission at $z\\sim3.4$, were described in Cowie \\& Hu\\markcite{smitty1} 1998 (hereafter, Paper I), and yielded a number of candidate high-$z$ galaxies similar to the redshift $z\\sim4.55$ Ly$\\alpha$-emitting galaxies found by Hu \\& McMahon\\markcite{br2237} (1996) and emitters at $z\\sim 2.4$ (Pascarelle et al.\\markcite{pasc96} 1996; Francis\\markcite{fra97} et al.\\ 1997) in targeted searches. In Paper I, we showed that the use of color selection on emission-line selected objects of high equivalent width ($> 100$ \\AA) picked out objects with continuum colors similar to those of color--selected Lyman break galaxies with measured redshifts of $z\\sim3.4$. They also recovered the one field object whose previously measured redshift placed Ly$\\alpha$ within the filter bandpass. However, the emission-line galaxies selected by their high equivalent width also comprised objects with very faint continuua, that would have fallen below the magnitude threshold of current Lyman break surveys, and also included two objects that could not be detected in Keck imaging of the optical continuum ($1\\ \\sigma$ $B=27.8$, $V=27.5$, and $I=25.8$; $W_{\\lambda}>400$ \\AA). In the present Letter we first present spectroscopic followups for the narrowband candidates of Paper I. In a small fraction of the cases, the spectrum shows both Ly$\\alpha$ and \\civ\\ $\\lambda$ 1550 \\AA, confirming the redshift identification but suggesting AGN-like properties. However, the majority of the spectra show only a single strong emission line. The absence of other detectable features, in combination with the high equivalent width of the selected candidates, identifes the single line as redshifted Ly$\\alpha$ emission, and argues that the equivalent width criterion ($W_{\\lambda}\\gg 100$ \\AA) can, in fact, be used as a good diagnostic of high-$z$ Ly$\\alpha$-emitting galaxies. We then (\\S3) present results of a second deep narrowband search with a 78 \\AA\\ bandpass $\\lambda \\sim6741$ \\AA\\ filter (Ly$\\alpha$ at $z\\sim 4.54$) and followup spectroscopy, that confirmed two Ly$\\alpha$-emitting galaxies at this higher redshift. In \\S4, we describe a very deep blank-field spectroscopic search (6 hr LRIS integration covering $\\lambda\\lambda\\,\\sim5000-10000$ \\AA) that yielded four emitters at redshifts 3.04 -- 5.64. Finally (\\S5), the data on emission-line objects from the imaging surveys in the two redshift intervals are combined with various spectroscopic surveys at lower redshift to show the evolution of the emission-line fluxes with redshift. We argue that the Ly$\\alpha$-emitting objects are significant contributors to the integrated star formation of the galaxy population throughout the $z=3 - 6$ redshift range, and that the integrated star formation rate of the Ly$\\alpha$ selected objects is flat, or possibly rising through this redshift range, with a value greater than 0.01 $M_{\\odot}$ Mpc$^{-3}$ for $q_0 = 0.5$ and $H_0 = 65\\ {\\rm km}\\ {\\rm s}^{-1}\\ {\\rm Mpc}^{-1}$. At the highest redshifts, most of the star formation may be occurring in objects of this class. ", "conclusions": "Since resonant scattering enhances the effects of extinction, it is harder to convert the Ly$\\alpha$ emission into a massive star formation rate than it is for line luminosity diagnostics such as H$\\alpha$ and \\oii\\ 3727. For the present calculation, we assume that extinction may be neglected in computing the required massive star formation rates, which then constitute a minimum estimate. However, upward corrections to this value are unlikely to be larger than a factor of two, since the observed rest-frame equivalent widths lie in the $100-200$ \\AA\\ range --- close to the maximum values that are obtainable from ionization by a massive star population (Charlot \\& Fall\\markcite{charl93} 1993). Then, assuming case B recombination, we have $L($Ly$\\alpha$) = 8.7\\,$L($H$\\alpha$) (Brocklehurst\\markcite{brock} 1971), which using Kennicutt's\\markcite{kenn83} (1983) translation of $\\dot{M}$ from H$\\alpha$ luminosity, gives $\\dot{M}=(L($Ly$\\alpha$)/$10^{42}$ ergs s$^{-1}$) $M_{\\odot}$ yr$^{-1}$. In order to cross-calibrate to \\oii\\ fluxes at lower redshift we assume $f$(H$\\alpha$+\\nii) = $1.25\\,f$(\\oii) based on the mean values of the ratio in both the Gallego et al.\\markcite{gal95} (1995) and Hawaii Deep Survey (Cowie et al.\\markcite{large_sample} 1996) samples. We also assume $f$(H$\\alpha$+\\nii) = $1.33\\,f$(H$\\alpha$) (Kennicutt\\markcite{kenn83} 1983). In Figure~\\ref{fig:5} we compare the range of Ly$\\alpha$ luminosities to the range of line luminosities in lower redshift objects. The plot shows the quantity $z^2\\,f$ vs redshift, where H$\\alpha$+\\nii, \\oii, and Ly$\\alpha$ fluxes have been converted to H$\\alpha$ fluxes using the relationships above, and we have restricted ourselves to the imaging data in which the fluxes of the Ly$\\alpha$ are well determined. The comparison objects are drawn from the surveys of Gallego et al.\\markcite{gal_95} 1995 ({\\it filled boxes}), Songaila et al.\\markcite{ksurvey_3} 1994 ({\\it pluses}), and the Hawaii $B=25$ sample [{\\it open boxes\\/} (H$\\alpha$+\\nii), {\\it triangles\\/} (\\oii), and {\\it circles\\/} (Ly$\\alpha$)]. The solid ($q_0=0.5$) and dashed ($q_0=0.02$) lines on Fig.~\\ref{fig:5} show the fluxes corresponding to stellar mass production rates of 10 $M_{\\odot}$ yr$^{-1}$, 1 $M_{\\odot}$ yr$^{-1}$, and 0.1 $M_{\\odot}$ yr$^{-1}$ for $H_0 = 65\\ {\\rm km}\\ {\\rm s}^{-1}\\ {\\rm Mpc}^{-1}$. Maximum formation rates locally are around a few $M_{\\odot}$ yr$^{-1}$, rising to values of just over 10 $M_{\\odot}$ yr$^{-1}$ above $z=0.6$. The Ly$\\alpha$ fluxes at the higher redshifts are then consistent with this value or just slightly smaller depending on the extinction correction. The CADIS results (Thommes et al.\\markcite{cadis} 1998) would lie a factor of several times higher in flux than the two Hawaii filter samples, but both Keck spectroscopy and repeat Fabry-P\\'erot observations have disproved the original $z\\sim5.7$ candidate selection (Meisenheimer\\markcite{meisen} 1998). The minimum integrated star formation rates at $z=3.4$ and $z=4.5$ are 0.006 $M_{\\odot}$ Mpc$^{-3}$ yr$^{-1}$ and 0.01 $M_{\\odot}$ Mpc$^{-3}$ yr$^{-1}$ respectively for $q_0=0.5$ where the first value is slightly smaller than that given in Paper I since AGN-like objects are excluded. Both values are lower limits calculated in the absence of extinction and the $z=4.5$ value is based on a single field. Within the substantial uncertainties of the as yet small samples, the results suggest that the star formation rates in the strong emission line population are constant or may possibly be increasing with redshift from $z=3 - 6$, in contrast to color-based samples where the rate is declining at higher redshifts (Madau et al.\\markcite{madau96} 1996, \\markcite{madau98} 1998). This is consistent with the broad expectation that as we move to higher redshifts and earlier stages of galaxy formation, an increasingly larger fraction of the star formation should be in strong Ly$\\alpha$ emitters that correspond to the youngest galaxies." }, "9803/astro-ph9803227_arXiv.txt": { "abstract": "Black holes are by definition {\\it black}, and therefore cannot be directly observed by using electromagnetic radiations. Convincing identification of black holes must necessarily depend on the identification of a very specially behaving matter and radiation which surround them. A major problem in this subject of black hole astrophysics is to quantify the behaviour of matter and radiation close to the horizon. In this review, the subject of black hole accretion and outflow is systematically developed. It is shown that both the stationary as well as the non-stationary properties of the observed spectra could be generally understood by these solutions. It is suggested that the solutions of radiative hydrodynamic equations may produce clear spectral signatures of black holes. Other circumstantial evidences of black holes, both in the galactic centers as well as in binary systems, are also presented. ", "introduction": "Stellar mass black holes are the end products of stars. After the fuel is exhausted inside a normal star, the core collapses and the supernova explosion occurs. If the mass of the core is lower than, say, $\\sim 3M_\\odot$, the object formed at the center may be a neutron star. Otherwise, it is a black hole. Therefore, some of the compact binary systems should contain black holes. Similarly, core collapse in the proto-galactic phase could also produce supermassive black holes ($M \\sim 10^6$ to $10^9M_\\odot$). In spiral galaxies, the central black holes are less massive (say, $10^{6-7}\\ M_\\odot$), while in elliptical galaxies the central black holes are more massive (say, a few times $10^{8-9}\\ M_\\odot$). Astrophysical community generally believes that the black holes should exist because of the solid foundation of the theory of general relativity which predicts them. The problem remains that of identification. Black holes do not emit anything except Hawking radiation, which, for any typical mass of the astrophysical black holes is so cold (typically $60$ nano Kelvin for a solar mass black hole, and goes down inversely with increase in mass) that it would be entirely masked by the much hotter microwave background radiation. Classically, black holes are point-like with infinite density and are surrounded by an imaginary one-way membrane called `event horizon' of radius $R_g=2GM_{BH}/c^2$. Here, $G$ and $c$ are gravitational constant and velocity of light respectively, $M_{BH}$ is the mass of the black hole. $R_g$ is known as the Schwarzschild radius and is roughly equal to $30$ kilometers for a $10\\ M_\\odot$ black hole. For a maximally rotating (Kerr) black hole, the radius is half as small. Surrounding matter and radiation are pulled by the black hole only to disappear inside never to be observed again. Not even light, what to talk about matter, can escape to distant observers from regions within the horizon, making it {\\it impossible} to detect a black hole through direct observations. A positive identification must therefore rely on indirect and circumstantial evidences. In fact, the problem of identification of black holes boils down to the identification of surrounding matter which may behave in a `funny' way. We shall quantify the degree of `funniness' as we go along. In this {\\it review}, we discuss how a black hole could be identified. We first present elementary properties of the spacetime around a black hole and compare them with those of a Newtonian star. We discuss in great length the properties of the global solutions of equations which govern the behaviour of matter. We then show that the observations in the last couple of decades do agree with these properties. Towards the end we make a comparative study of methodologies of black hole detection and present our judgment on the best way to detect black holes. ", "conclusions": "That black holes, which represent the end product of massive stars and star clusters, must exist somewhere in this universe is beyond any doubt. The issues discussed in this review were: whether they are in principle detectable, how to detect them and whether they have been detected. It seems that a few cases at least they {\\it have been detected}. If the observations of Genzel et al. [96, 125] is correct, then the mass of the central $0.1pc$ region of our galaxy would be $2.5-3.2 \\times 10^6 M_\\odot$ and the corresponding mass density would be $6.5 \\times 10^9 M_\\odot /pc^3$, the highest measured concentration so far. The water mega-maser measurement of the nucleus of NGC4258 within $0.1pc$ has the central mass of $4 \\times 10^7 M_\\odot$ and corresponding mass density is $6.5 \\times 10^9 M_\\odot /pc^3$. The central mass of M87 from the estimation of Keplerian and non-Keplerian components is $ \\sim 4 \\times 10^9 M_\\odot$ and the corresponding mass density is $2.0 \\times 10^7 M_\\odot /pc^3$. Although, Cyg X-1 is the most studied black hole candidate so far, its mass function is very low. Its confirmation as a black hole comes from its spectral features, especially the weak power-law slope of the bulk motion Comptonization in its soft state. The only candidates with mass function higher than, say, $3 M_\\odot$, are $GRS1124-683$, $GRO J1655-40$, $H 1705-250$, $GS 2000+25$ and $GS2023+338$ and are possible stellar mass black holes. With the improvements of the future observational techniques, one needs to focus on more detailed predictions of the advective disks, such as variation of the solution topology with specific energy, or equivalently, accretion rate. With the emergence of gravitational wave astronomy, the wave signals from galactic centers should be detectable. The proposal presented in Section (5.3) would for the first time correlate the distortions of the gravitational wave signals with those from the spectral signatures. Together they would not only verify black holes, they may also become the strongest test of general relativity to date. \\newpage \\centerline {Reference}" }, "9803/astro-ph9803157_arXiv.txt": { "abstract": "The effects which star cluster concentration and binarity have on observable parameters, that characterise the dynamical state of a population of stars after their birth aggregate dissolves, are investigated. To this end, the correlations between ejection velocity, binary proportion, mean system mass, binary orbital period and mass ratio are quantified for simulated aggregates. These consist of a few hundred~low-mass binary and single stars, and have half-mass radii in the range~2.5 to 0.08~pc. The primordial binary-star population has a period distribution similar to that observed in Taurus-Auriga for pre-main sequence binaries. The findings presented here are useful for interpreting correlations between relative locations and proper motions, binary properties and masses of young stellar systems within and surrounding star forming regions, and of stellar systems escaping from Galactic clusters. For the low-concentration binary-rich aggregates, the proportion of binaries decreases monotonically as a function of increasing ejection velocity after aggregate dissolution, as expected. However, this is not the case for initially highly concentrated binary-rich aggregates. The reason for this difference is the interplay between the disruption of binary systems and the initial depth of the potential well from which the stellar systems escape. After aggregate dissolution, a slowly expanding remnant population remains. It can have a high binary proportion (80~per cent) with a high mean system mass, or a low binary proportion (less than about 20~per cent) with a low mean system mass, if it was born in a low- or a high-concentration aggregate, respectively. It follows that adjacent regions on the sky near some star-forming clouds can have young populations with different binary proportions and different mass functions, even if the binary proportion at birth and the initial mass function (IMF) were the same. Binary systems that are ejected from the aggregate tend to be massive, and their mass ratio tends to be biased towards higher values. The mean system mass is approximately independent of ejection velocity between~2 and~30~km/s. Dynamical ejection from binary-rich aggregates adds, within~10~Myr, relatively massive systems to regions as far as~300~pc from active star-forming centres. Long-period systems cannot survive accelerations to high velocities. The present experiments show that a long-period ($>10^4$~d) binary system with a large velocity ($>30$~km/s) cannot be ejected from an aggregate. If such young systems exist, then they will have been born in high-velocity clouds. ", "introduction": "\\label{sec:intro} \\noindent Stellar systems (i.e. single or multiple stars) form in groups. The dynamical processes within these alter the properties of the young systems when they leave the site where they formed. The dynamical properties of a stellar system are its mass (i.e. luminosity if age is known), the multiplicity and the orbital parameters if it is a multiple system. The distribution of velocities of young stars emanating from star-forming centres (i.e. the kinematical signature of star formation) will also be affected by the dynamical interactions within the young groups. Both, the distribution of {\\it dynamical properties} and the {\\it kinematical signature of star formation} bear an imprint of the dynamical configuration at the time when the stellar group was born. Star formation in Taurus-Auriga gave birth to aggregates with sizes of roughly 0.5--1~pc consisting of about 20--50 stars. It is now well established that most stars form in binary systems in Taurus-Auriga (e.g. K\\\"ohler \\& Leinert 1998). The same appears to hold true in other star-forming regions (Ghez et al. 1997). Embedded clusters may also have a binary proportion that is higher than in the Galactic field (Padgett, Strom \\& Ghez 1997). In the Trapezium cluster, which is a very dense embedded cluster and probably less than 1~Myr old, Prosser et al. (1994) find a binary proportion that is at least as large as in the Galactic field. In the cluster core, Petr et al. (1998) observe, for low-mass stars, a binary proportion similar to the Galactic field, and smaller by about a factor of three than the binary proportion in Taurus-Auriga. These findings are particularly interesting, because binary destruction is expected to be efficient in such an environment. A review of pre-main sequence binary stars, and their relation to Galactic field systems, is provided by Mathieu (1994, see also Kroupa 1995a, Simon et al. 1995). Young Galactic clusters also contain binary systems. The particularly well studied Pleiades and Praesepe clusters have binary proportions of 40--50~per cent, for systems of spectral type earlier than K0 (Raboud \\& Mermilliod 1998a, 1998b). There exists thus evidence that the formation of binary systems may be by far the dominant star-formation mode in both loose groups and highly concentrated embedded clusters, some of which may evolve to bound Galactic clusters. The term {\\it aggregates} is used henceforth to mean loose groups or embedded clusters of more than 10~stars. If stars form predominantly in aggregates of binary systems, then the kinetic energy distribution after aggregate dissolution should be enhanced at high energies, when compared to dissolved aggregates of single stars, because binary star binding energy can transform into kinetic energy (Heggie 1975, Hills 1975, Hut 1983). Large accelerations are destructive to binary systems, so that the proportion of binaries should be a decreasing function of increasing final kinetic energy. Additionally, different initial aggregate concentrations lead to different final binary proportions and kinematical signatures, as will be shown here. This is also true under the extreme assumption that {\\it all} stars always form in binary systems with the same initial dynamical properties. This dynamical mechanism of producing variations in binary proportion and associated dynamical properties stands in contrast to a possible variation of these parameters determined by the star-formation process. Durisen \\& Sterzik (1994) make the interesting point that the binary proportion may be smaller in molecular clouds with a higher temperature than in lower-temperature clouds. It is important to study the signatures that arise from purely dynamical interactions in stellar groups, for a comparison with outcomes from usually less well-understood alternative scenarios. In a study of the large-scale distribution of young stars around active star-forming regions, Sterzik \\& Durisen (1995) find that the dynamical decay of small stellar groups can lead to sufficiently large velocities to populate large areas on the sky with young stars, so that these need not have formed near their observed location. They find that special initial dynamical configurations of the stars (e.g. cold thin strings) lead to enhanced production of ejected stars. Initial decay of cold sub-groups within larger complexes also has this effect (Aarseth \\& Hills 1972), and scattering of proto-stars on cloud clumps during an even earlier dynamical phase may likewise eject very young low-mass stars (Gorti \\& Bhatt 1996). However, the number of ejected stars cannot account for the observed number of widely distributed young stars (Feigelson 1996). The evolution of circum-stellar discs around stars ejected from small stellar groups is studied by Armitage \\& Clarke (1997), and McDonald \\& Clarke (1995) show that the presence of circum-stellar material in small proto-stellar groups increases the number of binaries formed and randomises the mass-ratio distribution. That the number of dynamically ejected stars is increased significantly in binary-rich stellar aggregates, when compared to clusters consisting initially only of single stars, is shown by Kroupa (1995c). These simulations show that a mass-ratio distribution produced by randomly associating masses from the IMF, decays to the observed distribution for G-dwarf binaries, if most stars form in aggregates similar to observed embedded clusters. Also, initially more concentrated aggregates produce more stars with a high ejection velocity, the maximum of which increases with decreasing cluster radius. De la Fuente Marcos (1997) investigates the dependence on cluster richness, and finds that the mean ejection velocity increases for more initially populous clusters. Ejection velocities larger than a few hundred~km/s can be achieved in young star clusters containing massive primordial binaries (Leonard \\& Duncan 1990). This may explain the location of OB stars far from active star-forming sites. Leonard (1991) finds, on the basis of many scattering experiments, that the maximum ejection velocity is of the order of the escape velocity from the stellar surface of the most massive star. If its mass is $60\\,M_\\odot$, then a similar star can attain an ejection velocity of up to 700~km/s. A low-mass star may find itself fleeing with a velocity of up to 1400~km/s, after a surface-grazing encounter with such a star. A critical discussion of the possible origin of runaway OB stars is provided by Leonard (1995). He stresses that collisions of two stars during binary-binary interactions can produce runaway OB stars with very similar properties as in the alternative scenario, in which such stars result from a supernova explosion in close binary systems. An interesting and insightful discussion of the implications of the binary properties of runaway OB stars for the dynamical configuration of massive stars at birth is to be found in Clarke \\& Pringle (1992). In this paper, the correlations between stellar velocity, system mass and binary proportion that arise from aggregates with different initial concentration and consisting initially either of 400~single stars or of 200~binary systems, is studied. The resulting correlations are useful for interpreting the properties and distribution of young stars near and in star forming regions (see for example Brandner et al. 1996, Feigelson 1996, Frink et al. 1997). In Section~\\ref{sec:method} the assumptions, simulations and definitions are described. The results are presented in Section~\\ref{sec:results}, and Section~\\ref{sec:conclusions} contains the conclusions. ", "conclusions": "\\label{sec:conclusions} \\noindent The correlations between ejection velocity and the proportion of binaries, as well as their orbital parameters, have been quantified for a range of initial dynamical configurations. The correlations are useful in the study of stellar systems that are apparently ejected from Galactic clusters (see e.g. Frink et al. 1997), some of which are known to be rich in binaries (e.g. Raboud \\& Mermilliod 1998a, 1998b). Observed ejected binaries should show correlations as presented in Figs.~\\ref{fig:orbit1}, \\ref{fig:orbit2}, ~\\ref{fig:mass1} and~\\ref{fig:mass2}. The results for the binary-rich aggregates modelled here are also relevant for an understanding of the large-scale distribution of young stars, because most stars appear to form in aggregates with a high binary proportion. Additionally, the correlations contain information about the dynamical configuration at birth. For binary-rich aggregates containing a few hundred stars the following correlations result: (i) more tightly clustered aggregates lead to more stellar systems having larger ejection velocities and a smaller overall binary proportion, (ii) the large population of primordial binaries leads to a significantly enhanced number of systems with high-ejection velocities compared to single-star aggregates, (iii) systems with high ejection velocities have a significantly reduced binary proportion, (iv) binary stars with high ejection velocities have short-period orbits, and tend to be more massive with a mass-ratio biased towards unity, (v) the average system mass as a function of ejection velocity is defined above about 2~km/s by stochastic close encounters, so that systems more massive than $0.5\\,M_\\odot$ with high ejection velocities occur, and (vi) aggregates with $R_{0.5}\\le0.25$~pc lead to a complex dependence of the resulting binary proportion on velocity, whereas a stellar population emerging from less concentrated aggregates shows a monotonic decrease of the binary proportion with increasing velocity. For aggregates of a few hundred single stars one obtains: (i) more tightly clustered aggregates lead to an increased number of stellar systems with larger ejection velocities (but significantly less so than in the binary-rich aggregates), and an enhanced overall binary proportion that remains significantly below the observed binary proportion in the Galactic field, (ii) the binary proportion increases with ejection velocity, (iii) is as (iv) above, and (iv) is as (v) above. Remnant unbound young populations take long to disperse because they have a small velocity dispersion. The binary proportion and mean system mass (and thus the inferred IMF) of such a remnant population, sensitively depends on the initial dynamical configuration of the binary-rich birth aggregate. After emerging from the birth aggregate, the distribution of velocities of a young stellar population changes with time in the gravitational potential of the nearby molecular clouds. A substantial proportion of emerging stars is likely to remain bound to the parent molecular cloud until it ceases to exist. These findings are important for interpreting the spatial distribution, kinematics and binarity of young stars within and surrounding star-forming regions. Molecular clouds, in which stars form preferentially in dense embedded binary-rich clusters, should have an enhanced halo population of ejected and relatively binary poor ($f\\approx0.25$) young stellar systems. Also, young but binary-depleted groups of stars can be misinterpreted to be evidence for an environmental dependency of the binary-formation mechanism. For example, in fig.~6 of Brandner et al. (1996), the region US-B has more binaries than the region US-A, which also contains many more B~stars than US-B. The presence of B~stars suggests that the stars in US-A may have formed in dense embedded clusters. The stars seen in US-A would then constitute the $v\\simless1$~km/s remnant population for $R_{0.5}\\simless0.25$~pc (Fig.~\\ref{fig:binprop1}). Given the results of the present study, it is suggested that such a difference in binary proportion between two regions may be due to different initial dynamical configurations, and need not imply a dependence of the binary proportion on the star-forming environment. Important for the interpretation of the large-scale distribution of young stars surrounding star forming sites is the realisation that relatively massive systems are ejected with relatively large velocity (2--30~km/s, Fig.~\\ref{fig:binprop1}), which is a point also stressed by Sterzik \\& Durisen (1995). The X-ray surveys are flux limited and detect the massive stars (Wichmann et al. 1996), the presence of which around star-forming regions may be a natural consequence of the processes studied here. However, if some young binary systems are found to have orbital periods that place them above the dashed lines in the right panels of Figs.~\\ref{fig:orbit1}--\\ref{fig:orbit3}, then this would support the suggestion by Feigelson (1996), that some star-formation occurs in small high-velocity clouds." }, "9803/astro-ph9803196.txt": { "abstract": "} \\nc{\\eab}{ \\noindent The explanation of the observed galactic magnetic fields may require the existence of a primordial magnetic field. Such a field may arise during the early cosmological phase transitions, or because of other particle physics related phenomena in the very early universe reviewed here. The turbulent evolution of the initial, randomly fluctuating microscopic field to a large-scale macroscopic field can be described in terms of a shell model, which provides an approximation to the complete magnetohydrodynamics. The results indicate that there is an inverse cascade of magnetic energy whereby the coherence of the magnetic field is increased by many orders of magnitude. Cosmological seed fields roughly of the order of $10^{-20}$ G at the scale of protogalaxy, as required by the dynamo explanation of galactic magnetic fields, thus seem plausible. ", "introduction": "%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Apart from the baryon number and the spectrum of energy density fluctuations, the physical processes that took place in the very early universe do not have many consequences that could still be directly detectable today. Most observables have been washed away by the thermal bath of the pre-recombination era. One possibility, which has recently received increased attention, is offered by the large-scale magnetic fields observed in a number of galaxies, in galactic halos, and in clusters of galaxies \\cite{observe,becketal}. The astrophysical mechanism responsible for the origin of the galactic magnetic fields is not understood. Usually one postulates a small seed field, which can then be either enhanced by the compression of the protogalaxy, and/or exponentially amplified by the turbulent fluid motion as in the dynamo theory \\cite{dynamo}. The exciting possibility is that the seed field could be truly primordial \\cite{kulsrud}, in which case cosmic magnetic fields could provide direct information about the very early universe. Early magnetic fields could then play an important role in particle cosmology by modifying the dispersion or clustering properties of various particles. One particular example is the fate of the neutrino: because of their magnetic moments, Dirac neutrinos propagating in the background of a magnetic field would be subject to a spin flip \\cite{eers}, so that a left-handed neutrino can be turned into a right-handed neutrino, giving rise to an extra effective neutrino degree of freedom and thereby affecting primordial nucleosynthesis. Dark matter particles could also be sensitive to the presence of a magnetic field. For instance, axions couple to magnetic fields, but perhaps surprisingly, it can be shown that despite the coupling, cold axion oscillations are not much affected by the presence of a primordial magnetic field \\cite{jarkkoax}. The issue at hand is then: is it possible that primordial magnetic fields of significant strength exist? To answer this, first one has to find a mechanism in the early universe which is able to produce a large enough magnetic field. There are various proposals, a number of which are based on the early cosmological phase transitions, which are discussed in Sect 3. The second problem is to explain how the initial field, which is expected to be random as it is created by microphysics and having correlation lengths typical to microphysics, can grow up to be coherent enough at large length scales. This is a problem in magnetohydrodynamics which is discussed in Sect 4. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % ", "conclusions": "%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Explaining the galactic magnetic fields in terms of microphysical processes that took place when the universe was only ten billionth of a second old is a daunting task, which is not made easier by the complicated evolution of the magnetic field as it is twisted and tangled by the flow of plasma. It is nevertheless encouraging that mechanisms for generating primordial magnetic fields of suitable size exist, and in particular those based on the early cosmological phase transitions discussed in Sect. 3 look promising. At the same time the fact that there are so many possibilities tends to underline our ignorance of the details of the subsequent evolution of the magnetic field. The step from microphysics to macroscopic fields is a difficult one because of the very large magnetic Reynolds number of the early universe. However, different considerations, both analytic approximations, 2d simulations, as well as the full-fledged shell model computations which can account for turbulence, seem to point to the existence of an inverse cascade of magnetic energy. Moreover, as discussed in Sect. 4.3, the inverse cascade is obtained also in the presence of a large plasma viscosity. Therefore the primordial origin of the galactig magnetic fields is quite possible. Much theoretical work remains to be done, though. At the same time it is very important that progress is made on the observational front. In particular, measuring or setting a stringent limit on the intergalactic field, which could be possible in the near future as indicated in Sect. 2.3, would provide the testing ground for all theoretical scenarios. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%" }, "9803/astro-ph9803143_arXiv.txt": { "abstract": "s{We describe two new -- {\\it stochastic-geometrical} -- methods to obtain reliable velocity field statistics from N-body simulations and from any general density and velocity fluctuation field sampled at a discrete set of locations. These methods, the {\\it Voronoi tessellation method} and {\\it Delaunay tessellation method}, are based on the use of the Voronoi and Delaunay tessellations of the point distribution defined by the locations at which the velocity field is sampled. Adjusting themselves automatically to the density of sampling points, they represent the optimal estimator for volume-averaged quantities. They are therefore particularly suited for checking the validity of the predictions of quasi-linear analytical density and velocity field perturbation theory through the results of N-body simulations of structure formation. We illustrate the subsequent succesfull application of the two methods to estimate the bias-independent value of $\\Omega$ in the N-body simulations on the basis of the predictions of perturbation theory for the $\\Omega$-dependence of the moments and PDF of the velocity divergence in gravitational instability structure formation scenarios with Gaussian initial conditions. We will also shortly discuss practical and complicating issues involved in the obvious extension of the Voronoi and Delaunay method to the analysis of observational samples of galaxy peculiar velocities.} \\vskip 1.0cm ", "introduction": "The study of the large-scale cosmic velocity field is a very promising and crucial area for the understanding of structure formation. The cosmic velocity field is in particularly interesting because of its close relation to the underlying field of mass fluctuations. Indeed, on these large and (quasi)-linear scales the acceleration, and therefore the velocity, of any object is expected to have an exclusively gravitational origin so that it should be independent of its nature, whether it concerns a dark matter particle or a bright galaxy. Moreover, in the linear regime the generic gravitational instability scenario of structure formation predicts that at every location in the Universe the local velocity is related to the local acceleration, and hence the local mass density fluctuation field, through the same universal function of the cosmic density parameter $\\Omega$ (Peebles 1980), $f(\\Omega) \\ \\propto \\ \\Omega^{0.6}$. Because linear theory provides a good description on scales exceeding a few Megaparsec, the use of this straightforward relation implies the possibility of a simple inversion of the measured velocity field into a field that is directly proportional to the field of local mass density fluctuations. Such a procedure can then be invoked to infer the value of $\\Omega$, through a comparison of the resulting field with the field of mass density fluctuations in the same region. However, such a determination of $\\Omega$ may be contrived as the estimate of the mass density fluctuation field on the basis of the observed galaxy distribution may offer a biased view of the underlying mass distribution. By lack of a complete and self-consistent physical theory of galaxy formation, the commonly adopted approach is to make the simplifying assumption that the galaxy density $\\delta_g$ and the mass density $\\delta$ are related via a linear bias factor $b$, $\\delta_g\\,=\\,b\\,\\delta$. The comparison between the observed galaxy density fluctuation field and the local cosmic velocity field will therefore yield an estimate of the ratio $\\beta\\,=\\,{f(\\Omega)/b} \\, \\approx\\, {\\Omega^{0.6}/b}\\,.$. However, while numerous studies have yielded estimates of $\\beta$ in the range $\\beta \\approx 0.5-1.2$ (see Dekel 1994, Strauss \\& Willick 1995, for compilations of results), it has proven very cumbersome to subsequently disentangle the contribution of $\\Omega$ and $b$ to the quantity $\\beta$. In fact, it turns out to be impossible within procedures based on the linearity of the analysed velocity field. ", "conclusions": "The main incentive for developing the Delaunay and Voronoi method is provided by the wish to be able to infer a bias-independent value of $\\Omega$ through comparison of the velocity statistics obtained from the discrete point sample with those of analytical distributions. In figure 3 we show the PDFs of the velocity divergence $\\theta$ that were numerically determined by the Delaunay method for a range of N-body simulations, each with a different value of $\\Omega$. The solid curve shows the corresponding analytical distribution function $p(\\theta)$, for the cosmic epoch with the same dispersion $\\sigma_{\\theta}$. For contrast, each of the four frames also contains the dashed curve for the PDF in an Einstein-de Sitter universe with the same value of $\\sigma_{\\theta}$. Evidently, the Delaunay method is highly succesfull in reproducing the correct statistical distribution, and via the relations between the moments of the PDF we indeed obtain very good estimates of $\\Omega$. While figure 3 illustrates the potential power of the tessellation methods, we are obviously motivated to apply them to more practical situations and hence more cumbersome cases where selection and sampling effects and sampling errors are of crucial influence. In particular we hope to be able to develop a formalism capable of dealing with observational catalogues of peculiar velocities of galaxies. In previous work (Bernardeau \\& Van de Weygaert 1996, Bernardeau et al. 1997) we already adressed the issue of diluted samples. In those cases both methods yielded encouraging results. However, the true world will present problems ranging from the fact that one can measure galaxy velocities only along the line of sight to complicated selection effects like differential Malmquist bias (see e.g. Bertschinger et al. 1990, Dekel, Bertschinger \\& Faber 1990). Work on these issues is in progress, but they obviously provide a considerable complication." }, "9803/astro-ph9803233_arXiv.txt": { "abstract": "Parallax data from the Hipparcos mission allow the direct distance to open clusters to be compared with the distance inferred from main sequence (MS) fitting. There are surprising differences between the two distance measurements, which indicate either the need for changes in the cluster compositions or reddening, underlying problems with the technique of main sequence fitting, or systematic errors in the Hipparcos parallaxes at the 1 mas level. We examine the different possibilities, focusing on MS fitting in both metallicity-sensitive \\bv\\ and metallicity-insensitive $V-I$ for five well-studied systems (the Hyades, Pleiades, $\\alpha$ Per, Praesepe, and Coma Ber). The Hipparcos distances to the Hyades and $\\alpha$ Per are within 1 $\\sigma$ of the MS fitting distance in \\bv\\ and $V-I$, while the Hipparcos distances to Coma Ber and the Pleiades are in disagreement with the MS fitting distance at more than the 3 $\\sigma$ level. There are two Hipparcos measurements of the distance to Praesepe; one is in good agreement with the MS fitting distance and the other disagrees at the 2 $\\sigma$ level. The distance estimates from the different colors are in conflict with one another for Coma but in agreement for the Pleiades. Changes in the relative cluster metal abundances, age related effects, helium, and reddening are shown to be unlikely to explain the puzzling behavior of the Pleiades. We present evidence for spatially dependent systematic errors at the 1 mas level in the parallaxes of Pleiades stars. The implications of this result are discussed. ", "introduction": "Main sequence fitting is a basic tool used in the study of star clusters; the principle behind it is also used to estimate distances to field main sequence (MS) stars. The Hipparcos mission (ESA 1997) has provided parallaxes for a number of open cluster stars, which permits a direct determination of the distances to the open clusters which can be compared with distances obtained from MS fitting. There are surprising differences between distances obtained with these two methods; in this paper we explore possible explanations for them. MS fitting relies upon the Vogt-Russell theorem: the location of a star in the HR diagram is uniquely specified by its mass, composition, and age. This implies that we can infer the distance of a given cluster by comparing the apparent magnitudes of cluster stars with the absolute magnitudes of stars with known composition and distance. There are several possible approaches. Unevolved lower MS field stars with known distances or a cluster (such as the Hyades) of known distance can be used to construct an empirical MS. The distance to the cluster is inferred from the vertical shift needed to line up the cluster MS with the empirical MS. Clusters can also be compared with theoretical isochrones calibrated on the Sun; the latter method requires a color calibration which relates the model effective temperatures to the observed colors. Most nearby open clusters are close to the Sun in metal abundance, which minimizes uncertainties in the distance scale from variations in composition. There is also a large database of fundamental effective temperature measurements for stars near the solar [Fe/H], so the color calibrations should be relatively reliable. The nearby open clusters also have been extensively studied for membership, photometry, abundances, and reddening. For all of these reasons the open cluster distance scale has not been regarded as controversial, and evidence that MS fitting yields incorrect distances could have significant astrophysical importance. The Hipparcos mission has resulted in a large increase in the number of open cluster stars with measured parallaxes. This data allows the distance scale inferred from MS fitting to be compared with the distance scale inferred from trigonometric parallaxes. The recently announced Hipparcos determination of the mean parallax of the Pleiades cluster gives the result $8.61 \\pm 0.23$ milliarcsec \\markcite{vh97a} (van Leeuwen \\& Hansen Ruiz 1997a). This corresponds to a distance of $116\\pm3$ pc, or a distance modulus of $5.32\\pm0.06$ magnitude. Traditional determinations of the Pleiades distance (e.g., \\markcite{vb84} VandenBerg \\& Bridges 1984; \\markcite{s93} Soderblom et al. 1993), comparing the cluster's main sequence to that of nearby stars, lead to a distance modulus of about 5.6 mag ($d \\sim 130$ pc; $\\pi \\sim 7.7$ mas). Thus the Hipparcos parallax, being almost 1 mas larger than expected, suggests that the Pleiades cluster stars are systematically $\\sim 0.3$ magnitude fainter than we have thought up to now. Parallaxes for stars in other clusters have also been measured, and the results are compared with those obtained from MS fitting in Table 1 (data taken from \\markcite{phip97} Perryman et al. 1997, \\markcite{mhip97} Mermilliod et al. 1997, \\markcite{rhip97} Robichon et al. 1997). The standard reddening for the clusters is also indicated, along with a notation about whether or not differential reddening is present. The second column lists the cluster [Fe/H] values from Boesgaard \\& Friel (1990) and Friel \\& Boesgaard (1992); we have adopted their abundance scale for the clusters in the present study (see Section 4). Mermilliod et al. 1997 and Robichon et al. 1997 concluded that there is no simple explanation for the discrepancies between the MS fitting and Hipparcos distances, and that all of the possible classes of solutions appeared unsatisfactory. \\begin{deluxetable}{lcccccc} \\tablenum{1} \\tablecaption{Open Cluster Parameters} \\tablehead{ \\colhead{Cluster} & [Fe/H] & \\colhead{$m-M$} & \\colhead{$(m-M)_o$} & \\colhead{$(m-M)_o$} & \\colhead{$(m-M)_o$} & \\colhead{$E(B-V)$}\\\\ \\colhead{ } & & \\colhead{Apparent} & \\colhead{Lynga} & \\colhead{Hipparcos} & \\colhead{This paper} & \\colhead{mag}} \\startdata Hyades & $+0.13$ & 3.01 & 3.01 & 3.33${\\pm}0.01$ & 3.34${\\pm}0.04$ & 0.00\\nl Coma Ber & $-0.07$ & 4.49 & 4.49 & 4.73${\\pm}0.04$ & 4.54${\\pm}0.04$ & 0.00\\nl Pleiades & $-0.03$ & 5.61 & 5.48 & 5.33${\\pm}0.06$ & 5.60${\\pm}0.04$ & 0.04\\nl IC 2602 & & 6.02 & 5.89 & 5.84${\\pm}0.07$ & \\nodata & 0.04\\nl IC 2391 & & 5.96 & 5.92 & 5.83${\\pm}0.08$ & \\nodata & 0.01\\nl Praesepe & $+0.04$ & 5.99 & 5.99 & 6.24${\\pm}0.12$ & 6.16${\\pm}0.05$ & 0.00\\nl ${\\alpha}$ Per & $-0.05$ & 6.36 & 6.07 & 6.33${\\pm}0.09$ & 6.23${\\pm}0.06$ & 0.10\\tablenotemark{a}\\nl Blanco 1 & & 6.97 & 6.90 & 7.01${\\pm}0.26$ & \\nodata & 0.02\\nl IC 4756 & & 8.58 & 7.94 & 7.30${\\pm}0.19$ & \\nodata & 0.20\\tablenotemark{a}\\nl NGC 6475 & & 7.08 & 6.89 & 7.32${\\pm}0.19$ & \\nodata & 0.06\\nl NGC 6633 & & 8.01 & 7.47 & 7.32${\\pm}0.34$ & \\nodata & 0.17\\tablenotemark{a}\\nl Stock 2 & & 8.62 & 7.41 & 7.50${\\pm}0.32$ & \\nodata & 0.38\\tablenotemark{a}\\nl NGC 2516 & & 8.49 & 8.07 & 7.71${\\pm}0.15$ & \\nodata & 0.13\\nl NGC 3532 & & 8.53 & 8.40 & 8.10${\\pm}0.36$ & \\nodata & 0.04\\nl \\enddata \\tablenotetext{a}{Variable reddening} \\end{deluxetable} We note that a second calculation of the distance to Praesepe has been performed by \\markcite{vh97b} van Leeuwen \\& Hansen Ruiz (1997b), and they find a distance modulus of 6.49$\\pm$0.15 - in disagreement both with MS fitting and the Mermilliod et al. Hipparcos distance. For the purposes of this paper we have adopted the Mermilliod distance; if we were to adopt the VH97b distance to the cluster we would have to add Praesepe to the list of clusters with a significant (2 $\\sigma$) discrepancy between the MS fitting and Hipparcos distance scales. The first column of distance moduli in Table 1 lists the values cited as ``Lynga'' by Mermilliod et al. (1997) and Robichon et al. (1997). We note that these are {\\it apparent} distance moduli, needing considerable (up to 1.2 mag) corrections for extinction, and cannot be directly compared with the distance moduli $(m-M)_o$ calculated from the Hipparcos parallaxes. The second column in Table 1 lists the distance moduli which correspond to the cluster distances given in Lynga's (1987) Catalogue. These distances come from a variety of sources, are still scaled to a Hyades distance modulus of 3.01 mag, and need corrections for each clusters metallicity. One motivation for our study is to place MS fitting distances for open clusters on a consistent scale. In a paper in preparation, we have found that the MS fitting distances to some of the more distant open clusters are substantially different from the Lynga distances and in marked disagreement with the Hipparcos parallax distances. A second question is the precision of MS fitting estimates; we will show that accuracy at the 0.05 mag level is possible for well-studied systems. Our results for the clusters studied in this paper are in the fourth column. Discrepancies between the MS fitting distances and the Hipparcos distances could arise from several sources. As indicated above, one possibility is that the MS fitting distances need to be rederived on a consistent scale. Another possibility is that some of the basic properties of well-studied open clusters, such as composition, age, or reddening, need to be revised. If neither of these possibilities can reconcile the distance scales, then we are left with one of two important conclusions : either there are fundamental problems with MS fitting or there are unrecognized systematic errors in the Hipparcos parallaxes themselves. These issues are important for other questions as well. For example, recent proposed revisions to the globular cluster distance and age scales, based on Hipparcos parallaxes of subdwarfs, rely on the same MS fitting technique that gives rise to the puzzling distances to open clusters (\\markcite{r97} Reid 1997; \\markcite{g97} Gratton et al. 1997; \\markcite{c98} Chaboyer et al. 1998; but see also \\markcite{pmtv98} Pont et al. 1998). In this paper we address the essential issues raised above. The Pleiades, Praesepe, and $\\alpha$ Per are well-suited for a more detailed examination. There is good membership information and multicolor photometry for all three; $\\alpha$ Per is a system with an age comparable to that of the Pleiades (50 Myr vs. 100 Myr) and therefore it provides a test of age-related effects. We have also examined the Coma Ber star cluster, which has a low quoted error for its Hipparcos distance. In a companion paper (Soderblom et al. 1998) we have searched for field stars with accurate parallaxes and anomalous positions in the HR diagram. We begin by describing the theoretical models which we use and the open cluster data in section 2. In section 3 we begin with a comparison of the Pleiades, Praesepe, and $\\alpha$ Per in different colors. We then use the Hyades cluster to test the zero-point of our distance scale, check on the shape of the isochrones in the observational color-magnitude diagram, and to determine the sensitivity of distance estimates in different colors to changes in metal abundance. We then derive distance modulus estimates at both solar [Fe/H] and the individual abundances inferred from high-resolution spectroscopy for the Pleiades, Coma Ber, Praesepe, and $\\alpha$ Per using several different methods and both \\bv\\ and $V-I$. The Pleiades and Coma Ber are found to be in disagreement with the Hipparcos distance scale. We discuss the sensitivity of our results to age, composition, and reddening in section 4, and present evidence that the Hipparcos parallaxes may contain small-scale ($\\sim$1 deg) systematic effects $\\sim$1 mas in size, large enough to cause the Pleiades parallax discrepancy. Our conclusions are in section 5. ", "conclusions": "The results of Section 3 indicate that it is the Hipparcos distance to the Pleiades which is in the most serious conflict with MS fitting. In all of the other systems except Coma Ber, MS fitting in different colors yields distance results that are consistent with one another, normal helium, and [Fe/H] values from high resolution spectroscopy. Coma Ber may have an equally serious disagreement, but the unusual behavior of the cluster in $V-I$ suggests that other problems may be contributing to the discrepancy for it. We therefore examine in turn the various possible mechanisms that could reconcile the cluster distance scales for the Pleiades; in all cases we believe that they cannot do so. In a companion paper we show that the same conclusions result from an examination of nearby field stars \\markcite{s98} (Soderblom et al. 1998). We then proceed to an analysis of the Hipparcos parallaxes for the Pleiades, and show that there are indications of possible systematic errors that could be the origin of the discrepancy. The calculations that we have presented are standard stellar models. We have therefore not included physical processes such as gravitational settling, rotational mixing, magnetic fields, internal gravity waves, or mass loss, which are surely present. There are strong reasons for believing that these nonstandard effects will not influence the distance scale, although they could be potentially important for other issues. The single most important reason is the youth of the clusters that we have examined; detailed nonstandard calculations predict little, if any, effect for ages as young as the Pleiades. In addition, any such process would have to affect stars with a wide range in masses to a similar extent and be different among different clusters to explain the pattern that we see. Gravitational settling is minimal in young systems such as the Pleiades, and the degree to which helium and heavy elements sink depends strongly on the convection zone depth and thus the stellar mass. For example, helium and heavy element diffusion are a 10\\% fractional effect in the Sun, which is almost 50 times older than the Pleiades. The observed cluster lithium abundances require a mild envelope mixing process, and models with rotational mixing that are consistent with the lithium data predict little or no deep mixing (Pinsonneault 1997). In addition, the observed range in rotation rates in clusters is large, and any extra mixing would produce a spread in MS properties rather than a uniform shift in the distance estimates. Other physical processes could affect the results, but they are still subject to a variety of observational constraints which make a large effect unlikely. We have compared different standard model calculations, and the zero-point offset is small (0.01-0.03 mag for stars between 5600 and 7000 K, for example). The systematic errors in the standard model distance estimates is therefore also too small to explain the results that we have obtained. We now discuss age, composition, and reddening effects. \\subsection{Age and Stellar Activity} It is well-known that many young stars are heavily spotted; this could influence the color-temperature relationship and therefore the distance estimates for young systems such as the Pleiades and $\\alpha$ Per. In Figures 1 and 2 we compared these two clusters at the Hipparcos distances in our two colors; the Pleiades is clearly anomalous with respect to $\\alpha$ Per if the Hipparcos distance scale is adopted. Since $\\alpha$ Per is younger and has a larger population of rapid rotators, if anything $\\alpha$ Per should be more anomalous than the Pleiades if our MS fitting age estimates were in error because of activity. We note that similar conclusions can be obtained by comparing young and old field stars (Soderblom et al. 1998). The narrow width of the Pleiades MS also indicates that a wide range in stellar activity does not produce a significant effect on the color-temperature relationship. For all of these reasons we reject the idea that youth is responsible for the difference between the distance estimates. Another possibility is that activity could be influencing the Pleiades [Fe/H], which has been derived from LTE model atmospheres. If such a phenomenon were at play, it might lead to derived abundances being a function of line strength due to the direct effect of activity on the stronger lines formed at smaller depths in the photosphere. We have a number of high resolution spectra of Pleiades members that was originally obtained to study lithium abundances. We have analyzed the \\ion{Fe}{1} data in the cool Pleiades dwarfs and find no such [Fe/H]-line strength correlation. This does not exclude such a real correlation, though, given the influence of damping, which is adjusted to enforce such a lack of correlation. To the extent that our damping assumptions seem quite reasonable compared to numerous other fine spectroscopic analyses, and are consistently applied in both the stellar and solar analyses to yield line-by-line [Fe/H] values, the analysis suggests any such trends are not substantial. Regardless, any systematic error in the inferred mean [Fe/H] is greatly mitigated by the fact that the damping adjustments enforce consistency with the weaker lines, which are formed at deeper depths, and thus presumably are more immune from the direct effects of chromospheric activity. Activity in very young stars can manifest itself in the form of an effective veiling continuum. Such behavior would presumably weaken the line absorption, thus leading to {\\it underestimated} line strengths and, hence, abundances. Detailed NLTE line formation calculations to determine how the active Pleiades dwarfs' Fe and other metal abundances might be affected by activity, spots, convective flows, {\\it etc.\\/} would be of interest, but are unlikely to produce large errors for the reasons discussed above. \\subsection{Heavy Metals} \\subsubsection{The Cluster [Fe/H] Scale} Homogeneous Fe abundances are available for the Pleiades, Praesepe, and $\\alpha$ Per from the work of Boesgaard and collaborators. Independent modern fine analyses of these clusters (and a few others) by other investigators are available for comparison with their work. All the studies considered here derive self-consistent solar Fe abundances with which the stellar values are normalized. Such a careful differential procedure can greatly reduce errors introduced by varying assumptions concerning the solar Fe abundance, model atmospheres, $gf$ values, {\\it etc}. \\markcite{bbr88} Boesgaard {\\it et al.\\/}~(1988) determine a mean Pleiades iron abundance of [Fe/H]$=-0.03$ from analysis of 17 F stars. The mean star-by-star reddening they use is essentially identical to the value we have adopted. \\markcite{b88}Boesgaard (1989) determined a ``best'' Pleiades abundance by analyzing new data for 8 Pleiads; the result was [Fe/H]$=+0.02$. Boesgaard \\& Friel (1990) used new data for 12 of the same stars in Boesgaard {\\it et al.\\/} to find a mean [Fe/H]$=-0.03$. The single datum standard deviation in all these studies is ${\\sim}0.07$ dex. The 1${\\sigma}$ level error in the mean is 0.02-0.03 dex, so the internal statistical uncertainties appear to be small. \\markcite{ccc88}Cayrel {\\it et al.\\/}~(1988) derive a mean Pleiades [Fe/H] of $+0.13$ from analysis of four Pleiades dwarfs, three of which are significantly cooler (mid G) than the Boesgaard F stars. The standard deviation is 0.10 dex, which is somewhat smaller than their estimated individual errors; the error in the mean is ${\\sim}0.06$ dex. The ${\\sim}0.1$ dex offset between the Cayrel and Boesgaard values is representative of uncertainties in reddening (which enters via photometric $T_{\\rm eff}$ determinations by Boesgaard), the $T_{\\rm eff}$ determinations (the Cayrel values are based on H$\\alpha$ profiles), and other details. The Cayrel result is consistent with \\markcite{e86}Eggen's (1986) inference from narrow band photometry that the Pleiades [Fe/H] is near the Hyades value In order to increase the sample of Pleiades stars with [Fe/H] determinations, some of us (\\markcite{k97}King {\\it et al.\\/}~1997) have used high quality Keck spectra of two slowly rotating very cool ($T_{\\rm eff}{\\sim}4500$ K) Pleiades dwarfs to derive Fe abundances. Our $T_{\\rm eff}$ values are spectroscopic determinations from balancing the abundances as a function of excitation potential, and the normalized abundances are derived by comparison with similarly analyzed solar data on a line-by-line basis. The mean abundance is [Fe/H]$=+0.06$, with estimated errors in the mean of perhaps 0.05 dex. While comparison of the different studies indicates there may be systematic errors at the 0.1 dex level, we regard this (dis)agreement to be quite satisfactory given the ${\\sim}2000$ K range in $T_{\\rm eff}$, the disparate sources of data, and distinct methods used to derive $T_{\\rm eff}$. While a slightly sub-solar Fe abundance is often assumed for the Pleiades based on the Boesgaard \\& Friel results, the totality of the high-resolution spectroscopic evidence may be more consistent with a slightly super-solar value; our photometric [Fe/H] is consistent with solar [Fe/H]. Therefore, if anything the data suggest a distance modulus estimate larger than our MS fitting value rather than smaller. Fe abundances for Praesepe F dwarfs have been derived by \\markcite{bb88} Boesgaard \\& Budge (1988), Boesgaard (1989), and Friel \\& Boesgaard (1992). The resulting values are $=+0.14$, $+0.10$, and $+0.05$, with star-to-star scatter of 0.06-0.07 dex, and mean uncertainties of 0.03-0.04 dex; again, the internal precision is good. The zero-reddening assumed in their $T_{\\rm eff}$ determinations is identical to our assumption. Other detailed studies of numerous Praesepe stars comparison are lacking. Analysis of the primary component of the Praesepe SB2 KW367, a mid-G star which is significantly cooler than the Boesgaard F stars, by \\markcite{kh96}King \\& Hiltgen (1996) yielded [Fe/H]$=+0.01$ with an uncertainty near 0.05 dex. Again, systematic errors at the 0.1 dex are indicated by this limited comparison. Combined with the above results, we see that [Fe/H] for Praesepe is 0.00-0.15 dex larger than for the Pleiades, with a preference for the lower middle of this range. The results inferred from MS fitting are consistent with the upper end of the range. Boesgaard {\\it et al.\\/}~(1988), Boesgaard (1989), and Boesgaard \\& Friel (1990) derived Fe abundances in $\\alpha$ Per F stars. The mean [Fe/H] values are $-0.02$, $+0.00$, and $-0.05$. The $\\alpha$ Per Fe abundance seems nearly identical to the Boesgaard Pleiades estimate. The star-to-star scatter in the larger $\\alpha$ Per samples is 0.08-0.09 dex; mean uncertainties are ${\\sim}0.04$ dex. The mean of the individual $\\alpha$ Per reddening values employed by Boesgaard is ${\\sim}0.03$ dex lower than the single value adopted here. This difference might require a 0.05-0.10 dex increase in [Fe/H] for consistency with our assumptions. \\markcite{bls88}Balachandran {\\it et al.\\/}~(1988) determined Fe abundances in a very wide range (F to K type) of $\\alpha$ Per stars. The mean abundance of the stars not considered by them to be non-members is [Fe/H]$=+0.04$ with a star-to-star scatter of 0.14 dex; the mean internal error is only 0.02 dex. Their assumed reddening is identical to our value. The results of Boesgaard {\\it et al.\\/}~and Balachandran {\\it et al.\\/}~agree to within 0.1 dex, but when adjustment is made for the slightly different reddening assumptions, the agreement is within a few hundredths of a dex if not exact. Our photometric [Fe/H] is slightly sub-solar, at the 0.01-0.02 dex level. It thus appears that the Fe abundance of $\\alpha$ Per is not significantly larger than for the Pleiades. In sum, internal errors in the Fe abundances of main sequence Pleiades, Praesepe, and $\\alpha$ Per stars derived from careful homogeneous analyses employing high quality data lead to uncertainties of only 0.05-0.10 dex in relative cluster abundances. We have seen that systematic effects due to errors in reddening, differences in the analysis methodology, {\\it etc.\\/} may approach 0.15 dex. These are small compared to the offset needed to explain the Hipparcos-based M$_V$ values for the Pleiades. Barring fundamental failure or incompleteness in our understanding of spectral line formation and stellar atmospheres, the extant data suggests that the Fe abundances of the Pleiades, Praesepe, and $\\alpha$ Per are within $\\sim0.10$ dex of each other. We might caution, however, that the abundances of other important atmospheric opacity contributors ({\\it e.g., Mg and Si}) are, unfortunately, unknown. \\subsubsection{Photometric Constraints and the Binary Distance to the Pleiades} There are other factors that make a large error in the Pleiades [Fe/H] unlikely. Colors that incorporate an infrared band are less sensitive to metallicity than \\bv. The figures in the previous section indicate clearly that the shift in the cluster distance modulus is the same for different color indices; the Pleiades must be intrinsically subluminous if the revised distance estimate is correct. The deviations from the high-resolution [Fe/H] values for the Pleiades are both large and inconsistent from color to color. The spectroscopic binary HD 23642 also provides a distance of $5.61\\pm0.26$ consistent with MS fitting, albeit with a large error \\markcite{gia95}(Giannuzzi 1995). \\subsection{CNO Abundances} Carbon, nitrogen, and oxygen can affect stellar structure in ways other elements do not; are they anomalous in the Pleiades? As part of his thesis, King (1993) examined the oxygen abundances of stars in several clusters over a broad range of age. The [O/H] for the Pleiades was found to be higher than for Praesepe (+0.29 and +0.02 respectively, with errors in the mean of 0.08 for both). However, the trustworthiness of abundances (such as these) derived from the high excitation 7774 \\AA \\ion{O}{1} lines is a matter of some debate. Besides possible large data and analysis differences between various studies (e.g. King \\& Hiltgen 1996), there may be significant abundance corrections due to non-LTE effects on line formation in stellar atmospheres (see Garcia Lopez et al. 1995). Unfortunately, systematic errors of 0.3 dex in the cluster O abundances derived from high-excitation lines remains plausible. In any case, the King results would act to make the Pleiades more metal-rich and therefore require a higher distance modulus estimate. Detailed abundance studies would be useful, but deviations from the solar mixture would need to be very large to have a significant impact on the luminosity of the MS. \\subsection{Helium} The initial solar helium abundance can be inferred from theoretical solar models by the requirement that the model have the solar luminosity at the age of the Sun. Modern evolution codes give estimates for the initial solar $Y$ in the range $0.26 - 0.28$; the best solar models of Bahcall, Pinsonneault, \\& Wasserburg (1995) had $Y = 0.272$ and $Y = 0.278$ with and without gravitational settling respectively. A comparison of theoretical stellar models with the Hipparcos main sequence of the Hyades by Perryman et al. (1997) yields $Y = 0.26\\pm0.02$; for comparison, the solar $Y$ in that study was 0.266 and the solar-scaled helium for the cluster would be 0.28. This agreement between the Sun and Hyades was anticipated and reinforces the notion that stars formed in the current epoch have similar helium abundances. Nevertheless, we consider what range of $Y$ would be needed to drop the Pleiades main sequence by 0.3 mag, and that value is about $Y = 0.37$. Such a high value of $Y$ for the Pleiades would imply a drastic revision of chemical evolution models and, by extension, would raise the possibility that other clusters might have similar anomalies. MS fitting would therefore require knowledge of both the metal and helium abundances; since helium can only be directly observed in young systems this would make MS fitting unreliable at the 0.3 magnitude level for the majority of clusters. We believe that this question is best answered by direct measurements of the helium abundance in HII regions and massive stars. We begin with a discussion of the literature on helium abundances; we have also obtained data on the relative helium abundances in the Pleiades and $\\alpha$ Per. Neither the field star data nor our Pleiades spectra are consistent with significant variations in the initial helium abundance from the solar value. Ignoring a deviant few percent of field stars, \\markcite{n74}Nissen's (1974) study revealed no intrinsic scatter in $Y$ greater than ${\\sim}10$\\% (compared to the 30-40\\% deviation required by the Pleiades stars) in nearby main-sequence field B stars. \\markcite{gl92} Gies \\& Lambert (1992) found helium abundances consistent with both the Sun and the Orion nebula for a sample of 35 B dwarfs; 4 B supergiants in that sample were found to have anomalously high helium abundances. There is evidence that evolutionary effects are responsible for helium enrichment in the most massive stars (see \\markcite{mc94}Maeder \\& Conti 1994, \\markcite{lu96}Lyubimkov 1996, \\markcite{pin97}Pinsonneault 1997 for reviews), so helium abundances from MS O stars and massive supergiants may not be reliable indicators of the initial $Y$. The B star field data and the Orion nebula abundances are therefore our best test for the range in helium abundance at solar metal abundance, and they are consistent with only small variations in the initial helium abundance. For Galactic clusters, however, the picture is less clear. \\markcite{ss69}Shipman \\& Strom (1969), \\markcite{ps73} Peterson \\& Shipman (1973), \\markcite{n76}Nissen (1976), and \\markcite{l77}Lyubimkov (1977) found evidence for 20\\%-30\\% variations in $Y$ among young associations, including some systems with significantly lower $Y$. Lyubimkov suggested an increasing He abundance with {\\it increasing\\/} age amongst the young clusters/associations studied, a conclusion not supported by the subsequent field star work of Gies \\& Lambert. \\markcite{p79}Patton (1979) determined He abundances of 60 stars in 8 young clusters and associations. She noted that her initial abundances displayed a range in $Y$ of about 25\\%, and that this could not be explained by the the usual error sources; she also called attention to a correspondence between He abundance and cluster age. However, Patton shows that binarity may be responsible for observed cluster-to-cluster He abundance dispersions, and the notably low He abundances (observed by others too) seen for a few stars within a given cluster/association. Eliminating {\\it suspected\\/} (but not positively identified) binary systems from her analysis results in cluster He abundances which are identical to within the uncertainties. This highlights the need for secure knowledge of very fundamental stellar parameters (e.g., binarity) before reliable He abundances can be derived. With this muddled picture of main sequence stellar He abundances, one may wonder if the Pleiades He abundance could be abnormal. Both the Pleiades and $\\alpha$ Per are young enough to have B stars, and their helium can be directly measured. The $Y$ values from Lyubimkov (1977) agree to within ${\\Delta}Y{\\sim}0.015$, which is well within the uncertainties; the Pleiades and $\\alpha$ Per $Y$ value is 0.04 larger than the corresponding field star value, but the uncertainties are comparable to this offset. \\markcite{kp86} Klochkova \\& Panchuk (1986) also derived B-star He abundances in both the Pleiades and $\\alpha$ Per. They claim to find no difference between the mean abundances that is larger than the uncertainties. However, this conclusion is not clear to us from the abundances listed in their Table II, which do demonstrate quite a very large difference. Unfortunately, only two Pleiades stars are included in the analysis. Therefore small number statistics and the possible effects of binarity make assessment of this difference quite difficult. We attempted a final comparison using the ``field'' stars from Nissen (1974). This sample includes four $\\alpha$ Per stars, and two stars (HR 5191 and 7121) which are suggested members of the purported Pleiades supercluster. The mean $Y$ value is only 0.03 larger for the Pleiades field stars than for the $\\alpha$ Per stars; the uncertainties are probably not much smaller than this difference. To investigate the possibility of a non-standard helium abundance in the Pleiades \\markcite{fk98} Fischer \\& King (1998) observed MS B stars in $\\alpha$ Per and Pleiades to differentially compare the helium abundances. Preliminary analysis of the lines strengths for six He lines suggests that the cluster He abundances are identical within an uncertainty of 15\\%. Any real difference appears to be in the opposite sense of what is needed to make the Pleiades underluminous: the Pleiades line strengths are, if anything, consistently smaller than the $\\alpha$ Per counterparts. \\subsection {Reddening and Systematic Errors in the Photometry} Reddening will tend to make a cluster MS fainter at a given color. If the reddening is increased the inferred distance modulus will therefore increase. The effect can be roughly estimated as follows : in the color interval that we are using for MS fitting the derivative of $M_V$ with respect to both \\bv\\ and $V-I$ is $\\sim$5. The extinction $A_V$=3.12$E(B-V)$ and $E(V-I)_K$=1.5$E(B-V)$. Adding these effects together an increase in $E(B-V)$ of 0.10 magnitudes would increase V at fixed \\bv\\ and fixed $V-I$ by 0.188 (0.5 mag from a shift of 0.1 in \\bv\\ - 0.312 mag from extinction) and 0.438 (0.75 mag from a shift of 0.15 in $V-I$ - 0.312 mag from extinction) magnitudes respectively. The relative distances inferred by the two colors can therefore be affected if the reddening is incorrect. In addition the [Fe/H] abundances derived for cluster stars are sensitive to T$_{eff}$, and an increased reddening would imply a higher [Fe/H] for a given equivalent width (therefore further increasing the distance modulus). Other colors, such as $R-I$, will be less reddening-sensitive. Neither the Hyades nor Praesepe show any evidence for reddening along the line of sight; increasing the reddening estimate for the Pleiades would worsen the discrepancy with the Hipparcos distance modulus estimate. Even changing $E(B-V)$ from 0.04 to 0 would only decrease the distance modulus by 0.08 magnitudes. The reddening estimates for the Pleiades have been derived for a wide range of masses and from different techniques; Crawford and Barnes used Stromgren photometry to estimate $A_V$ for B, A, and early F stars in the Pleiades and Praesepe, Prosser and Stauffer used M dwarfs in the same clusters, and Breger used polarization measurements in the Pleiades. We conclude that reddening is not a significant source of uncertainty in distance estimates for the Pleiades. Multicolor distance measurements of the type performed in this paper could be a useful check on the reddening for more heavily obscured systems. Another possibility is that systematic errors in the photometry could cause errors in the distance estimates. For the color range that we are considering, the slope of the MS is $\\sim$5; this would require a systematic error of 0.06 magnitudes in \\bv\\ to reconcile the Pleiades distance scales, which is unreasonably large. The size of the systematic errors can be constrained by comparing spectroscopic temperature estimates with those based upon colors. In the case of Coma Ber, for example, it appears that spectroscopic temperature estimates are in agreement with the \\bv\\ colors of F stars but not with the $V-I$ colors. We note that the slope of the MS in $V-I$ is steeper for F stars than for the cooler stars, and that systematic errors in the $V-I$ photometry might explain the puzzling behavior of Coma Ber. We have attempted whenever possible to rely upon a single source for photometry in a given color for a given cluster. Even in the case of the $V-I$ data we see no evidence of systematic differences between the location on the color magnitude diagram of stars with colors converted to the Cousins system from the Kron system and those converted to the Cousins system from the Johnson system. For the Pleiades, independent studies (Section 2.2) give consistent photometry for individual stars at the level of the quoted errors (0.01 - 0.02 mag). The 0.3 mag discrepancy between the Hipparcos and MS fitting distance distance modulii is much too large to be explained by systematic errors in the photometry. High-resolution spectroscopy of the Pleiades is consistent with the observed colors, and the reddening is small. For systems with higher reddening, however, care must be taken when converting between different photometric systems; the Johnson, Cousins, and Kron system I bands have different effective central wavelengths and therefore different reddening corrections. \\subsection{Systematic Errors in the Hipparcos Parallaxes} The final possibility is that the Hipparcos Pleiades parallaxes may contain previously undetected systematic errors. If the MS fitting result $m-M = 5.60$ does indeed give the correct Pleiades distance, then a systematic zero-point error would need to approach the 1 mas level to produce the discordance with the Hipparcos results. Such an error seems impossibly large, in view of the extensive tests \\markcite{ar95,ar97} (Arenou et al. 1995, 1997) demonstrating the global zero-point error of the Hipparcos parallaxes to be smaller than 0.1 mas. However, global tests have little power to reveal effects occurring on the small angular scale ($\\sim 1 \\deg$) of the Hipparcos spatial correlations (see below). Indeed, the Hipparcos parallaxes of stars in open clusters such as the Pleiades represent the first real opportunity to test for systematic effects on small angular scales. One might well argue that it would only be prudent to consider the Hipparcos cluster results as the first direct tests for small-scale zero-point errors, rather than as cluster distance measurements. The Hipparcos Pleiades parallax (van Leeuwen \\& Hansen Ruiz 1997a) is based on measurements of 54 cluster members, ranging in $V$ from 2.8 to 11.5 within $5\\deg$ of the cluster center, so it represents a fairly broad sampling of the cluster. Because Hipparcos observed widely separated ($\\sim 58\\deg$ apart) star fields simultaneously, the parallaxes are inherently on an absolute scale over the whole sky. Over small regions of the sky ($\\lesssim 2\\deg$), however, the astrometric results are positively correlated because neighboring stars (within the $0.9\\deg \\times \\ 0.9\\deg$ Hipparcos field of view) tended to be observed on the same great circles the satellite swept out over the sky (Lindegren 1988, 1989). A comprehensive discussion of the Hipparcos mission and data reductions is given in Volumes 1--3 of the Hipparcos Catalogue \\markcite{ESA97}(ESA 1997). The spatial correlations may significantly impact the astrometric results for star clusters, whose angular size is of the same order as the Hipparcos correlation scale. To account for this, \\markcite{vh97a} van Leeuwen \\& Hansen Ruiz (1997a) re-calculated the Pleiades mean parallax from the intermediate Hipparcos data. For this paper, one of us (R.B.H.) has re-examined the individual Pleiades parallaxes from the Hipparcos Catalogue. Moreover, besides the spatial correlations, there is a different type of correlation affecting the Hipparcos results -- the statistical correlations among the five astrometric parameters (position, proper motion, and parallax), arising from the imperfect distribution of Hipparcos observations on the sky over time. In classical parallax work (cf. \\markcite{va75} Vasilevskis 1975), the time distribution of observations over a star's parallactic ellipse is controlled to maximize the parallax factors and minimize the correlations between position, proper motion, and parallax. This is easy to achieve from the ground, but Hipparcos could not do this because of the limited span of observations and the pattern of scans of the sky, as explained in Section 3.2.4 (pp. 321-325) of the Hipparcos Introduction (ESA 1997, Vol. 1). Figures 3.2.42 to 3.2.61 of that work illustrate the patterns of the correlations over the sky; Figure 3.2.66 (p. 363) shows histograms of the 10 correlations. The RMS values are $\\sim 0.2$, and large areas of the sky show correlations averaging 0.4 or more in size. It must be emphasized that these correlations are substantially larger than would be considered acceptable in ground-based parallax observations. For parallax work, the most important correlation is $\\rho_\\alpha^\\pi\\ $, between parallax and right ascension (Field H20 in the Hipparcos Catalogue). This is because, over most of the sky, most of the extent of the parallactic ellipse is in right ascension. The Hipparcos $\\rho_\\alpha^\\pi$ correlation is shown in Fig. 3.2.44 of the Hipparcos Introduction. Large values of $\\rho_\\alpha^\\pi$ were caused in certain areas of the sky by the unfortunate circumstance of unequal observations on both sides of the Sun, as discussed on p. 325 of the Hipparcos Introduction. This happens to impact the Pleiades particularly badly. The mean value of $\\rho_\\alpha^\\pi$ near the Pleiades center is +0.4; this is at the 96th percentile in the histogram in Fig 3.2.66. The question this raises is whether this large correlation, caused by the time distribution of Hipparcos observations of the Pleiades stars, has any effect on the parallax values. We stress again that this is a different effect from the spatial correlation that exists because Hipparcos astrometric data over small ($\\sim 1 \\deg$) areas of the sky are not fully independent measurements. In Figure 19 we plot parallax vs. the correlation $\\rho_\\alpha^\\pi$ for 49 Pleiades members verified by proper motion, radial velocity, and position in the color-magnitude diagram. (Mermilliod et al's 51 stars and van Leeuwen et al's 54 are virtually the same set as these; we rejected several additional stars on account of problems noted in Fields H30 and H59 of the Hipparcos Catalogue.) This plot shows several interesting things. The filled symbols are 12 bright ($V < 7$) stars within $\\sim 1 \\deg$ of the cluster center with correlations $\\rho_\\alpha^\\pi \\geq +0.34$ (the mean value for the whole sample). Due to the spatial correlation effect, these 12 stars all have nearly the same parallax (mean 8.86 mas, RMS dispersion 0.45 mas; $\\chi^2$ too small at the 0.995 significance level). Because Hipparcos' errors are smallest for bright stars, these stars carry much of the weight of the Pleiades parallax. There is a clear trend (slope) of parallax vs. $\\rho_\\alpha^\\pi$ correlation; a weighted least-squares solution gives a slope of $+3.04 \\pm 1.36$ mas per unit correlation. The solid line in Fig. 19 is this slope, run through the mean point (+0.34,+8.53). The dashed lines show $\\pm1\\sigma$ slopes. The intercept at zero correlation is $\\pi = 7.49 \\pm 0.50$ mas, quite consistent with the MS fitting distance. Figure 20 plots parallax vs. distance from the cluster center. The filled symbols are the same 12 bright stars with high $\\rho_\\alpha^\\pi$ as in Fig. 16. The open symbols are the 15 stars with $\\rho_\\alpha^\\pi < +0.25$, with no restriction on magnitude or distance. The two sets of stars barely overlap because the brightest stars in the Pleiades are highly concentrated to the cluster center. The low-correlation stars lie farther from the Pleiades center and show a much larger parallax scatter, reflecting (a) the larger errors for fainter stars and (b) the lack of spatial correlations on scales $\\gtrsim 1 \\deg$. Moreover, their mean parallax is smaller (reflecting the slope discussed above). For the 15 stars with $\\rho_\\alpha^\\pi < +0.25$, the weighted mean parallax is $7.46 \\pm 0.43$ mas. The RMS dispersion is 1.66 mas, consistent with the published parallax errors. This exercise is not intended to be a definitive re-determination of the Pleiades parallax; that would require going back to the intermediate Hipparcos data as per van Leeuwen et al (1997), and exploring the effects of both the $\\rho_\\alpha^\\pi$ and the spatial correlations at that level. However, it is quite clear that (a) small-angular-scale systematic effects at the 1 mas level are present in the Hipparcos Pleiades parallaxes; (b) these effects are related to the high values of the $\\rho_\\alpha^\\pi$ correlation near the cluster center; (c) the bright stars within $\\sim 1\\deg$ of the center, which carry most of the weight of the mean parallax, are the most severely affected; and (d) the stars with lower $\\rho_\\alpha^\\pi$ correlations, far enough ($\\gtrsim 1 \\deg$) from the center to be unaffected by the spatial correlation, have smaller parallaxes, consistent with the MS fitting distance. We also looked for effects of the $\\rho_\\alpha^\\pi$ correlation in the Hyades, Praesepe, $\\alpha$~Per, and Coma~Ber clusters. In Figures 21--24 we present the parallax vs. correlation plots for those clusters. The Hyades, Praesepe, and $\\alpha$~Per clusters also have large values of $\\rho_\\alpha^\\pi$, but the the slope (d$\\pi$/d$\\rho$) present in the Pleiades data does not occur in these clusters, where the MS fitting distances and the Hipparcos distances are in good agreement. The data for Coma~Ber do show a slope d$\\pi$/d$\\rho \\ = \\ -4.0 \\pm 2.1$ mas, but the range of $\\rho_\\alpha^\\pi$ is small, and the mean is near zero." }, "9803/astro-ph9803005_arXiv.txt": { "abstract": "We report on a mechanism which may lead to a spin-up of the surface of a rotating single star leaving the Hayashi line, which is much stronger than the spin-up expected from the mere contraction of the star. By analyzing rigidly rotating, convective stellar envelopes, we qualitatively work out the mechanism through which these envelopes may be spun up or down by mass loss through their lower or upper boundary, respectively. We find that the first case describes the situation in retreating convective envelopes, which tend to retain most of the angular momentum while becoming less massive, thereby increasing the specific angular momentum in the convection zone and thus in the layers close to the stellar surface. We explore the spin-up mechanism quantitatively in a stellar evolution calculation of a rotating $12\\,\\Msun$ star, which is found to be spun up to critical rotation after leaving the red supergiant branch. We discuss implications of this spin-up for the circumstellar matter around several types of stars, i.e., post-AGB stars, {\\Be} stars, pre-main sequence stars, and, in particular, the progenitor of Supernova 1987A. ", "introduction": "\\lSect{intro} The circumstellar matter around many stars shows a remarkable axial symmetry. Famous examples comprise Supernova~1987A (Plait et al. 1995; Burrows et al. 1995), the Homunculus nebula around $\\eta$~Carina and other nebulae around so called Luminous Blue Variables (Nota et al. 1995), and many planetary nebulae (Schwarz et al. 1992). A less spectacular example are {\\Be} stars, blue supergiants showing properties which might be well explained by a circumstellar disk (Gummersbach et al. 1995; Zickgraf et al. 1996). Many of these axisymmetric structures have been explained in terms of interacting winds of rotating stars (cf. Martin \\& Arnett 1995; Langer et al. 1998; Garc\\'{\\i}a-Segura et al. 1998), which may be axisymmetric when the stars rotate with a considerable fraction of the break-up rate (Ignace et al. 1996; Owocki et al. 1996). However, up to now only little information is available about the evolution of the surface rotational velocity of stars with time, in particular for their post-main sequence phases. Single stars which evolve into red giants or supergiants may be subject to a significant spin-down (Endal \\& Sofia 1979; Pinsonneault et al. 1991). Their radius increases strongly, and if the specific angular momentum were conserved in their surface layers (which may not be the case; see below) they would not only spin down but they would also evolve further away from critical rotation. Moreover, they may lose angular momentum through a stellar wind. Therefore, it may appear doubtful at first whether post-red giant or supergiant single stars can retain enough angular momentum to produce aspherical winds due to rotation. However, by investigating the evolution of rotating massive single stars, we found that red supergiants, when they evolve off the Hayashi line toward the blue part of the Hertzsprung-Russell (HR) diagram may spin up dramatically, much stronger than expected from local angular momentum conservation. In the next Section, we describe the spin-up mechanism and its critical ingredients. In Section~3 we present the results of evolutionary calculations for a rotating $12\\,\\Msun$ star, which provides a quantitative example for the spin-up. In Section~4 we discuss the relevance of our results for various types of stars, and we present our conclusions in Section~5. ", "conclusions": "\\lSect{con} In this paper, we discussed the effect of mass outflow through the inner or outer boundary of a rigidly rotating envelope on its rotation frequency. It causes a change of the specific angular momentum in the envelope and alters its rotation rate besides what results from contraction or expansion (cf. \\Fig{t-v}). For constant upper and lower boundaries of the envelope, which we found a good approximation for convective envelopes (cf. \\Fig{t-r-m}), a spin-down occurs for mass outflow through the upper boundary --- which corresponds, e.g., to the case of stellar wind mass loss from a convective envelope (cf. also Langer \\mbox{1998) ---,} while a spin-up results from mass outflow through the lower boundary (cf. \\Fig{panels}). The latter situation is found in evolutionary models of a rotating $12\\,\\Msun$ star at the transition from the Hayashi-line to the blue supergiant stage. The star increased its rotational velocity one order of magnitude above the velocity which would result in the case of local angular momentum conservation. It would have increased its rotational velocity even further if it would not have arrived at critical rotation (cf. \\Sect{res}, \\Fig{t-v}), with the consequence of strong mass and angular momentum loss. At this point, the specific angular momentum loss $\\Jdot/\\Mdot$ reached about $8\\,10^{19}\\,\\junit$ (cf. \\Fig{t-xx}). The geometry of circumstellar matter around stars which undergo a red $\\to$ blue transition may be strongly affected by the spin-up. We propose that this was the case for the progenitor of SN~1987A, the only star of which we know that it performed a red $\\to$ blue transition in the recent past. The blue supergiant in its neighborhood studied by Brandner et al. (1997), around which they found a ring nebula as well, is another candidate. Also, {\\Be}~stars may be related with the post-red supergiant spin-up (cf. \\Sect{blue-loop}). Furthermore, the spin-up mechanism studied in this paper may be relevant for bipolar outflows from central stars of proto-planetary nebulae (\\Sect{post-AGB}), from stars in the transition phase from the red supergiant stage to the Wolf-Rayet stage (\\Sect{RSG-WR}), and from pre-main sequence stars (\\Sect{pre-MS})." }, "9803/astro-ph9803280_arXiv.txt": { "abstract": "I present 1.5- and 8.4-GHz observations with all configurations of the NRAO VLA of the wide-angle tail source \\Ss{3C130}. The source has a pair of relatively symmetrical, well-collimated inner jets, one of which terminates in a compact hot spot. Archival {\\it ROSAT} PSPC data confirm that 3C\\,130's environment is a luminous cluster with little sign of sub-structure in the X-ray-emitting plasma. I compare the source to other wide-angle tail objects and discuss the properties of the class as a whole. None of the currently popular models is entirely satisfactory in accounting for the disruption of the jets in 3C\\,130. ", "introduction": "\\Ss{3C130} is a FRI radio source at redshift 0.109 (Spinrad \\etal\\ 1985). Its 178-MHz luminosity is $7.6 \\times 10^{25}$ W Hz$^{-1}$ sr$^{-1}$, slightly above the nominal FRI-FRII boundary of $\\sim 2 \\times 10^{25}$ W Hz$^{-1}$ sr$^{-1}$ (Fanaroff \\& Riley 1974, hereafter FR). Leahy (1985, 1993) and J\\\"agers and de Grijp (1985) present intermediate-resolution VLA maps of the central regions of the source, while J\\\"agers (1983) has a lower-resolution WSRT image which shows the whole source and its field; the source extends for $\\sim 1.5$ Mpc. Saripalli \\etal\\ (1996) present high-frequency maps made with the Effelsberg 100-m telescope. The host galaxy is classed as a DE2 by Wyndham (1966) and appears to lie in a cluster, although strong galactic reddening makes optical identification of the cluster members difficult. The {\\it Einstein} detection of extended X-ray emission (Miley \\etal\\ 1983), the near\\-by align\\-ed sources (J\\\"agers 1983) and the many mJy radio sources in the field at 1.5 GHz make it plausible that the object is the dominant member of a large cluster. Leahy (1985) also attempts to constrain the RM distribution of the source, but notes that it depolarizes rapidly (particularly in the S lobe) so that few good measurements are available; this could be taken as evidence for a dense magneto-ionic environment for the source (cf.\\ Hydra A, Taylor \\etal\\ 1990). \\label{definition} 3C\\,130 is a wide-angle tail (WAT) radio source. The term WAT has been used to describe many different types of object. Here I shall use it to refer to those FRI sources which are associated with central cluster galaxies (e.g.\\ Owen \\& Rudnick 1976) and have luminosities comparable to or exceeding the Fanaroff-Riley break between FRI and FRII. I shall follow Leahy (1993) in using the behaviour of the jets at the base as another defining feature. At high resolution one or two well-collimated jets [`strong-flavour' jets, by the classification of Leahy (1993)] are seen (e.g.\\ O'Donoghue, Owen \\& Eilek 1990), extending for some tens of kpc before broadening, often at a bright flare point, into the characteristic plumes or tails. These jets are very similar to the jets seen in FRII radio galaxies, and quite different from the behaviour of jets in more typical FRIs, where a collimated inner jet, if visible at all, decollimates rapidly (on scales of a few kpc at most) and comparatively smoothly into a bright `weak-flavour' jet with a large opening angle.\\footnote{There are a few exceptions to this behaviour; \\Ss{3C66B} (Hardcastle \\etal\\ 1996) does appear to show an inner `strong-flavour' jet and a bright knot at the base of the `weak-flavour' jet. But even here the transition from strong to weak flavours occurs on scales of $\\sim 1$ kpc.} WATs, according to this definition, never have a weak-flavour jet, but make the transition between strong-flavour jet and diffuse, bent tail in a single step. The requirement that WATs be central cluster galaxies excludes objects (e.g. \\Ss{3C171}, Blundell 1996, Hardcastle \\etal\\ 1997a; \\Ss{3C305}, Leahy 1997) where the `tails' are likely to be simply ordinary FRII lobes which have been disrupted by unusual host-galactic dynamics. The condition on jet behaviour allows us to exclude objects such as the twin sources in \\Ss{3C75} (Owen \\etal\\ 1985; Hardcastle 1996) which are associated with a dominant cluster galaxy and sometimes classed as WATs but whose inner jets are similar to those of typical powerful FRIs. Because of the requirements of this definition, wide-angle tail sources make up a small minority of the radio source population. For this reason, the detailed properties of their jets and tails have not been well studied, although a number have been imaged for studies of source dynamics (O'Donoghue \\etal\\ 1989). The only objects which have been the subject of detailed study in the radio are \\Ss{3C465} (Leahy 1984; Eilek \\etal\\ 1984) and \\Ss{3C218}, Hydra A (Taylor \\etal\\ 1990), although M87, Virgo A (e.g.\\ Biretta \\& Meisenheimer 1993) exhibits some of the properties of a WAT. In this paper I present multi-configuration, multi-frequency VLA observations of a further powerful WAT. Throughout this paper I use a cosmology in which $H_0 = 50{\\rm\\ km\\,s^{-1}\\,Mpc^{-1}}$ and $q_0 = 0$. At the distance of \\Ss{3C130}, one arcsecond is equivalent to a projected length of 2.72 kpc. B1950.0 co-ordinates are used throughout. ", "conclusions": "A compact hot spot is detected at the base of one plume of the WAT \\Ss{3C130}, and the jets are shown to have longitudinal magnetic field. The source is thus very like a classical double in some respects. The data support the model in which WATs are objects whose jets make the transition from super- to sub-sonic velocities in one step, rather than decelerating gradually, by showing a bright sub-kpc structure (comparable to those seen in classical double radio sources) associated with the termination of a jet. Archival {\\it ROSAT} PSPC observations of 3C\\,130 show it to lie in a luminous cluster with $kT \\sim 2.9$ keV. There is little sign of substructure in the X-ray, in contrast to many other WATs; this may be related to the nearly straight tails of 3C\\,130. The lack of strong substructure seems to be inconsistent with recent models for jet disruption in WATs." }, "9803/astro-ph9803249_arXiv.txt": { "abstract": "We predict the rate at which Gamma-Ray Burst (GRB) afterglows should be detected in supernova searches as a function of limiting flux. Although GRB afterglows are rarer than supernovae, they are detectable at greater distances because of their higher intrinsic luminosity. Assuming that GRBs trace the cosmic star formation history and that every GRB gives rise to a bright afterglow, we find that the average detection rate of supernovae and afterglows should be comparable at limiting magnitudes brighter than $K=18$. The actual rate of afterglows is expected to be somewhat lower since only a fraction of all $\\gamma$--ray selected GRBs were observed to have associated afterglows. However, the rate could also be higher if the initial $\\gamma$--ray emission from GRB sources is more beamed than their late afterglow emission. Hence, current and future supernova searches can place strong constraints on the afterglow appearance fraction and the initial beaming angle of GRB sources. ", "introduction": "Since their discovery in the late 1960's (Klebasadel et al. 1973) through early 1997, Gamma-Ray Bursts (GRBs) had defied all attempts to determine their distance scale conclusively. The Burst And Transient Source Experiment (BATSE) on board the Compton Gamma-Ray Observatory (GRO) showed that the burst population is highly isotropic (Meegan et al. 1993; Briggs et al. 1993), suggesting that bursts occur at cosmological distances or in an extended Galactic halo. Moreover, the cumulative number counts of faint bursts deviated from that of a uniform distribution of sources in Euclidean space and flattened at faint fluxes, consistent with the expected effect of a cosmological redshift (Fishman \\& Meegan 1995, and references therein). Last year, with the advent of the BeppoSAX satellite (Boella et al. 1997), it became possible to localize GRB sources to within an arcminute on a timescale of hours. Such fast, accurate localizations were quickly followed by the detection of delayed X-ray (Costa et al. 1997), optical (van Paradijs et al. 1997), and radio (Frail et al. 1997) counterparts to GRB sources. In particular, FeII and MgII absorption lines were detected at a redshift $z=0.835$ in the spectrum of the optical counterpart to GRB970508 (Metzger et al. 1997), demonstrating conclusively that this burst occurred at a cosmological distance with a redshift $z>0.835$. The isotropy of the burst population and the flattening of their number counts, taken in combination with the fact that the first confirmed redshift for an optical counterpart is high, provides strong evidence that GRB sources are located at cosmological distances. Most plausible GRB models involve either the collapse of a single massive star (e.g. Usov 1992; Woosley 1993; Paczy\\'nski 1998), or the coalescence of two compact objects -- two neutron stars or a neutron star and a black hole -- in a binary system (e.g. Paczy\\'nsky 1986; Eichler et al. 1989; Narayan et al. 1992; Mochkovitch et al. 1993; Rees 1997). Since the lifetime of these progenitors is short compared to the Hubble time at a redshift $z\\la 5$, the cosmic GRB rate should simply be proportional to the star formation rate at these redshifts, without any appreciable delay due to the finite progenitor lifetime. The cosmic rate of massive star formation rate has been determined from the $U$ and $B$-band luminosity density in Hubble Deep Field (Madau et al. 1996; Madau 1996; Madau, Pozzetti, \\& Dickinson 1997; Madau 1997). The inferred star formation rate $\\dot\\rho_{\\rm s}(z)$ can then be converted to a GRB explosion rate $R_{\\rm GRB}(z)$, based on the requirement that the latter would fit the observed number count distribution of $\\gamma$--ray selected GRBs (Wijers et al. 1997). Cosmological GRBs are at least $10^4$ times rarer than Type II supernovae (SNeII) -- possibly even $\\sim 10^6$ times rarer if GRBs occur primarily at high redshifts following the cosmic star formation history (Wijers et al. 1997). However, at peak luminosity, the GRB afterglows are $\\sim 10^3$--$10^4$ times brighter than SNeII. In Euclidean space, this would imply that GRBs are detected from a volume bigger by a factor $\\sim (10^{4})^{3/2}= 10^6$, roughly canceling out the factor by which they are rarer than supernovae. Hence we expect that at some relatively bright limiting flux, the rate of afterglow detections should become comparable to that of SN detections. Current and future supernova searches should provide information about the fraction of GRBs which produce detectable afterglows. The statistics of bursts in 1997 for which afterglows could have been identified implies that this fraction is of order tens of percent (e.g., Castro-Tirado 1998). On the other hand, there could also be a population of afterglows without a GRB precursor. This would occur if the source emits a jet from which the $\\gamma$--ray emission is more beamed than the subsequent optical afterglow radiation due to the deceleration of the jet by the ambient gas and the corresponding decline in its relativistic beaming with time (Rhoads 1997). A jet geometry would imply a higher rate of afterglow detections in supernova searches. In this {\\it Letter}, we predict the detection frequency of GRB afterglows as a function of limiting flux at various observed wavelengths, and compare this rate with the analogous predictions for SNe Type Ia and Type II at high redshifts. We assume throughout a flat, $\\Omega =1$, $\\Lambda =0$, cosmology, with a Hubble constant $H_0=50$ km s$^{-1}$ Mpc$^{-1}$. ", "conclusions": "" }, "9803/astro-ph9803139_arXiv.txt": { "abstract": "We analyse the population of bright star clusters in the interacting galaxy pair NGC 4038/39 detected with HST WFPC1 by Whitmore \\& Schweizer (1995). Making use of our spectrophotometric evolutionary synthesis models for various initial metallicities we derive the ages of these star clusters and calculate their future luminosity evolution. This allows us to compare their luminosity function ({\\bf LF}), evolved over a Hubble time, to LFs observed for the Milky Way's and other galaxies' star cluster systems. Since effective radii are difficult to determine due to crowding of the clusters, the shape of the LF after a Hubble time may help decide whether the young clusters are young globular clusters ({\\bf GC}) or rather open clusters/OB associations. We find an intriguing difference in the shapes of the LFs if we subdivide the cluster population into subsamples with small and large effective radii. While the LF for the extended clusters looks exponential, that for clusters with small effective radii clearly shows a turn-over brighter than the completeness limit. For other possible subdivisions as to luminosity or colour no comparable differences are found. Evolving, in a first step, the LF from a common mean age of the young clusters of 0.2 Gyr to an assumed age of 12 Gyr, the LF for the subsample of clusters with small effective radii seems compatible with a Gaussian GCLF with typical parameters M$_{\\rm V_0} = -7.1$ and $\\sigma (\\rm M_{\\rm V_0}) = 1.3$ except for some overpopulation of the faint bins. These faintest bins, however, are suspected to be subject to the strongest depopulation through effects of dynamical evolution not included in our models. We also follow the colour evolution of the young star clusters over a Hubble time and compare to observations on the Milky Way and other galaxies' GC systems. For an ongoing starburst like the one in the NGC 4038/39 system age spread effects among the young star cluster population may not be negligible. In a second step, we therefore account for age spread effects, instead of using a mean age for the young cluster population, and this drastically changes the time evolution of the LF, confirming Meurer's (1995) conjecture. We find that $-$ if age spread effects are properly accounted for $-$ the LF of the entire young star cluster population, and in particular that of the brighter subsample, after a Hubble time is in good agreement with the average Gauss-shaped LF of globular cluster systems having a turn-over at $\\langle {\\rm M_{V_0}} \\rangle = -7.1$ mag and $\\sigma({\\rm M_{V_0}}) = 1.3$ mag. The age distribution shows that the brightest globular clusters from the interacting galaxies' original population are also observed. They make up the bulk of the red subpopulation with (V$-$I)$_0 > 0.95$. Their effective radii do not significantly differ from those of the young star cluster population, neither on average nor in their distribution. We discuss the influence of metallicity, the effects of an inhomogeneous internal dust distribution, as well as the possible influence of internal $-$ through stellar mass loss $-$ and external dynamical effects on the secular evolution of the LF. Referring YSC luminosities to a uniform age and combining with model M/L, we recover the intrinsic mass distribution of the YSC system. It is Gaussian in shape to good approximation thus representing a quasi-equilibrium distribution that $-$ according to Vesperini's (1997) dynamical modelling for the Milky Way GC system $-$ will {\\bf not} be altered in shape over a Hubble time of dynamical evolution, allthough a substantial number of clusters will be destroyed. We briefly compare the young star cluster population of the Antennae to the older one in the merger remnant NGC 7252 and point out that the intercomparison of young cluster populations in an age sequence of interacting and merged galaxies may become an interesting approach to study in detail the role of external dynamical effects. ", "introduction": "From the fact that $-$ when normalised to the stellar mass of a galaxy $-$ the specific globular cluster ({\\bf GC}) frequency $T_{GC} := {N_{GC} \\over {M_{\\ast} / 10^9~M_{\\odot}}}$ is a factor of $\\sim 2$ higher in ellipticals than in spirals, Zepf \\& Ashman (1993) predict that if elliptical galaxies are formed from one major spiral $-$ spiral merger the number of GCs formed during the merger- induced starburst should be of the same order of magnitude as the number of GCs present in the progenitor galaxies. The high burst strengths and star formation ({\\bf SF}) efficiencies in massive gas-rich spiral $-$ spiral mergers and in IR-ultraluminous galaxies led to expect the formation of star clusters so tightly bound that they are able to survive as GCs (Fritze $-$ v. Alvensleben \\& Gerhard 1994). Fritze $-$ v. Alvensleben \\& Gerhard (1994) predicted the metallicity range of stars and star clusters formed in massive gas-rich (i.e. late type) spiral$-$spiral mergers on the basis of the ISM abundances of the progenitor galaxies to be $\\third ~{\\rm Z_{\\odot} \\lta Z \\lta Z_{\\odot}}$ or $-0.8 \\lta {\\rm [Fe/H]} \\lta -0.2$. In many interacting galaxies and merger remnants, bright blue knots have by now been observed (cf. e.g. Lutz 1991, Holtzman \\etal 1992, Whitmore \\etal 1993, Hunter \\etal 1994, O'Connell \\etal 1994, 1995, Conti \\& Vacca 1994, Borne 1996, Meurer \\etal 1995). These bright blue knots, of course, immediately raised the question as to their identity: are these Young Star Clusters ({\\bf YSC}) $-$ or, at least, some of them $-$ the progenitors of GCs? And, if the latter were true, how many of them are typically formed in a merger? How many will be able to survive in the tidal field of two massive interacting spirals? Can such a higher metallicity subpopulation be identified in GC systems (hereafter {\\bf GCS}) around merger remnants and perhaps even around normal ellipticals? Could the metallicity distribution of a GCS give information about the origin of its parent galaxy (cf. Zepf \\& Ashman 1993)? Or should all of these bright blue knots be open clusters/OB associations (van den Bergh 1995) most of which will disperse within few Gyr? The discussion of the nature of these YSCs is focussed on two aspects, their effective radii R$_{\\rm eff}$ and their luminosity function. In mergers at distances of the Antennae or NGC 7252, effective radii as measured on WFPC1 images are clearly overestimated. However, it has been shown that for YSC systems close enough the mean effective radii do readily fall within the range of GC radii (Meurer \\etal 1995). Our focus in this paper is the luminosity and colour evolution of the YSC population in the Antennae and, in particular, the future evolution of the YSC's LF. In a previous paper, we model the evolution of star clusters for different initial metallicities in terms of broad band colours and stellar metallicity indices. We find important colour differences for clusters of various metallicities, already at young ages, and showed that once the stellar metallicity is known, rather precise age dating becomes possible. Comparison with young star clusters in NGC 7252 (Whitmore \\etal 1993), the two brightest of which have spectroscopy available (Schweizer \\& Seitzer 1993), confirmed a metallicity of ${\\rm Z \\sim \\half Z_{\\odot}}$ predicted from our global starburst modelling in this Sc $-$ Sc merger remnant. The mean age of the young star cluster population was shown to agree well with the global burst age of $\\sim 1.3$ Gyr, and ages derived from solar metallicity models would differ by a factor $\\sim 2$ (see Fritze $-$ v. Alvensleben \\& Burkert 1995 for details). \\medskip\\noindent Observationally, the best case by now to study the LF of YSCs are the Antennae with more than 700 young star clusters detected by Whitmore \\& Schweizer (1995, hereafter {\\bf WS95}), a number large enough to allow for a statistical analysis. In this paper, we will examine the LF of the young star cluster system in the Antennae. It seems clear that not all bright knots in the NGC 4038/39 system with its still ongoing starburst will probably be GCs, in particular those with large effective radii R$_{\\rm eff}$ might rather be open clusters or associations. Therefore, after age dating the clusters in Sect. 2., we subdivide Whitmore \\& Schweizer's young star cluster sample into two subsamples containing the small knots and the more extended systems, respectively (Sect. 3.). In a first step, we assume a uniform age for the YSC population and we model the evolution of the YSCs' LF over a Hubble time and compare to LFs of the Milky Way's and other nearby galaxies' GCSs (Sect.4.). In an ongoing starburst like in the Antennae, the age spread among the YSCs may not be negligible (see also Meurer 1995). To examine the age spread effects on the LF we determe individual ages for all star clusters from their (V-I) colour and discuss the star clusters' age distribution in Sect. 5. We calculate the resulting individual fading for all clusters in Sect. 6. Alternative possibilities to subdivide the YSC sample and their consequences are discussed in Sect. 7. The of a young GCS may not only change by fading but also by dynamical effects as e.g. stellar mass loss within the cluster and/or tidal interaction of a cluster with the galactic potential. For GC populations in non-interacting galaxies, these effects were studied by Chernoff \\& Weinberg (1990), their results are largely confirmed by the independent and more realistic approach of Fukushige \\& Heggie (1995). In a recent paper Vesperini (1997) shows that in the Milky Way potential an initial log-normal mas distribution represents a quasi-equilibrium state that allows to preserve both its shape and parameters during a Hubble time of dynamical evolution, even though up to 70 \\% of the initial cluster population get disrupted. In case of the Antennae, i.e. in a still uncompleted merger with its gravitational potential being highly variable both in space and in time, however, external dynamical effects seem extremely difficult to model. Referring YSC luminosities to a common age allows to recover the mass function of the YSC system when combined with model M/L. We discuss the possible influence of dynamical effects in Sect. 8. and point out the possibility to observationally approach these dynamical effects by intercomparing star cluster populations in interacting galaxies and merger remnants of various ages. Sect. 9. summarizes our conclusions. The spatial distribution of the YSCs $-$ and of their properties as derived here $-$ will be discussed in a forthcoming paper. ", "conclusions": "Using our method of evolutionary synthesis for various metallicities we present a first analysis of WS95's WFPC1 data on bright star clusters in the ongoing merger-induced starburst in NGC 4038/39. Assuming a metallicity Z $\\sim 0.01$ on the basis of the progenitor spirals' ISM properties and applying a uniform reddening as given by WS95 we age-date the bright cluster population from their (V$-$I) colors and, as far as available, also from their (U$-$V). It turns out that in addition to a large population of young clusters with a mean age of $2 \\cdot 10^8$ yr (consistent with the dynamical time since pericenter) part of the original spirals' old GC population is also observed. A key question with far-reaching consequences as to the origin of elliptical galaxies is whether there are a significant fraction of young GCs among the YSC population. Two basic properties discriminate open clusters/OB associations from GCs in our Galaxy and others: the concentration parameter c = log (${\\rm R_T/R_{eff}}$) and the LF which, in contrast to that for an open cluster system, is Gaussian for {\\bf old} GCSs. Tidal radii and, consequently, concentration parameters not being accessible to observations in distant galaxies we examine the LFs of cluster subsamples with large and small effective radii. In a first step, using a common mean age for all young clusters and a corresponding uniform fading to an age of $\\sim 12$ Gyr we find that while the LF for extended clusters at 12 Gyr is definitely not Gaussian, that for the low R$_{{\\rm eff}}$ clusters may well contain a Gaussian (= GC) subcomponent together with a strong overpopulation of the faint bins, which themselves, however, might be expected to be severely depopulated over a Hubble time by dynamical effects not included in our models. Since for an ongoing starburst the age spread among YSCs may be of the same order as their ages, age spread effects are expected to reshape the LF. Clusters from the bright end tend to be younger on average and fade more than clusters from the faint end. We therefore, in a second step, model the individual fading consistent with individual ages of the YSCs as derived from their ${\\rm (V-I)}$ and ${\\rm (U-V)}$ colours, and we follow the LF changing its shape over a Hubble time. Surprisingly, accounting for these age spread effcets, we find the final LFs of large {\\bf and} small R$_{\\rm{eff}}$ cluster subsamples not to be significantly different any more. Instead, the LF of {\\bf all} YSCs evolved to a common age of 12 Gyr is well compatible with a ``normal'' GCLF. Its turn-over occurs at $\\langle {\\rm M_{V_0}} \\rangle \\sim -6.9$ mag, i.e. slightly fainter than the average value $\\langle {\\rm M_{V_0}} \\rangle \\sim -7.1$ mag for 16 galaxies. This difference is readily explained in terms of a higher metallicity of the secondary cluster population. The number of old GCs from the spiral progenitors is consistent with the number of bright GCs expected if the progenitors had GCSs similar to the ones in the Milky Way and M31. Strikingly, neither the mean nor the distribution of effective radii is significantly different for the old GC sample and for the YSC sample. On the basis of these WFPC1 data we tentatively conclude that the bulk of the YSC population detected in the Antennae might well be young GCs and that the open clusters/associations probably also present among the YSCs do not seem to systematically differ from young GCs in terms of R$_{\\rm{eff}}$. We are looking foreward to repeat this kind of analysis on WFPC2 data which may reach close to the old GCS's turn-over, reveal a number of fainter young objects, and will allow for more precise and definite conclusions. Dynamical effects that eventually might further reshape the LF over a Hubble time are discussed. Referring the YSCs' luminosities to a uniform age allows to recover the intrinsic mass function of the YSC system. This mass function seems to be log-normal which, according to Vesperini (1997), represents a quasi-equilibrium distribution that is going to be preserved in shape though not in number of clusters over a Hubble time of dynamical evolution. Dynamical effects, however, are extremely difficult to model in detail in an ongoing merger. Comparison of YSC populations in mergers/starbursts of various ages seems a promising tool in an attempt to understand these effects from an observational side. \\vskip 1 cm {\\sl Acknowledgements.} I am deeply indebted to B. Whitmore \\& F. Schweizer for valuable discussions, encouragement and for sending us their star cluster data in machine readable form. I am grateful to Ken Freeman, Tom Richtler, and Andreas Burkert for interesting discussions on dynamical aspects. I wish to thank Prof. Appenzeller and all the collegues from the Landessternwarte Heidelberg for their warm hospitality during a 3 months stay, when this projected was begun. My deep thanks go to the referee, G. Meurer, for his very detailed and constructive suggestions that greatly improved the paper. I gratefully acknowledge financial support from the SFB Galaxienentwicklung in Heidelberg and through a Habilitationsstipendium from the Deutsche Forschungsgemeinschaft under grant Fr 916/2-1 in G\\\"ottingen." }, "9803/astro-ph9803162_arXiv.txt": { "abstract": "We have established a model to systematically estimate the contribution of the mid-infrared emission features between 3 $\\mu$m and 11.6 $\\mu$m to the IRAS in-band fluxes, using the results of ISO PHT-S observation of 16 galaxies by Lu et al. (1997). The model is used to estimate more properly the $k$-corrections for calculating the restframe 12 and 25 $\\mu$m fluxes and luminosities of IRAS galaxies. We have studied the 12-25 $\\mu$m color-luminosity relation for a sample of galaxies selected at 25 $\\mu$m. The color is found to correlate well with the 25 $\\mu$m luminosity, the mid-infrared luminosity, and the ratio of far-infrared and the blue luminosities. The relations with the mid-infrared luminosities are more sensitive to different populations of galaxies, while a single relation of the 12-25 $\\mu$m color vs. the ratio of the far-infrared and the blue luminosities applies equally well to these different populations. The luminous and ultraluminous infrared galaxies have redder 12-25 $\\mu$m colors than those of the quasars. These relations provide powerful tools to differentiate different populations of galaxies. The local luminosity function at 12 $\\mu$m provides the basis for interpreting the results of deep mid-infrared surveys planned or in progress with ISO, WIRE and SIRTF. We have selected a sample of 668 galaxies from the IRAS Faint Source Survey flux-density limited at 200 mJy at 12 $\\mu$m. A 12 $\\mu$m local luminosity function is derived and, for the first time in the literature, effects of density variation in the local universe are considered and corrected in the calculation of the 12 $\\mu$m luminosity function. It is also found that the 12 $\\mu$m-selected sample are dominated by quasars and active galaxies, which therefore strongly affect the 12 $\\mu$m luminosity function at high luminosities. The ultraluminous infrared galaxies are relatively rare at 12 $\\mu$m comparing with a 25 $\\mu$m sample. ", "introduction": "\\label{sec:intro} The mid-infrared (MIR) spectral region is well-suited for studying starburst and ultraluminous galaxies. About 40\\% of the luminosity from starburst galaxies is radiated from 8-40 $\\mu$m (\\markcite{soi87}Soifer et al. 1987). Extinction effects are small, and infrared cirrus emission is reduced at these wavelengths relative to far-infrared bands. For a fixed telescope aperture, the spatial resolution is also higher at shorter wavelengths, and the confusion limit lies at higher redshifts. All the recent and near-future infrared space missions, such as the {\\it Infrared Space Observatory (ISO)}, the {\\it Wide-Field Infrared Explorer (WIRE)}, and the {\\it Space Infrared Telescope Facility (SIRTF)} will conduct surveys in mid-infrared bands. {\\it WIRE}, a Small Explorer mission due to launch in late 1998 (\\markcite{hac96}Hacking et al.\\ 1996; \\markcite{schemb96}Schember et al. 1996), will conduct a very deep survey at 12 and 24 $\\mu$m to study the evolution of starburst galaxies. To interpret the results of these surveys now in progress or soon to commence, it is necessary to better understand the mid-infrared properties of galaxies in the local Universe. One of the most important tools for extracting the rate and type of galaxy evolution from a mid-infrared survey is the faint source counts. A local mid-infrared luminosity function is the basis for calculating the mid-infrared faint source counts incorporating different evolutionary scenarios, and to extract the evolution by comparing with observations. Mid-infrared luminosity functions have been calculated at 12 and 25 $\\mu$m (Soifer \\& Neugebauer 1991) using a 60 $\\mu$m selected IRAS sample (Soifer et al. 1987). More recently, \\markcite{rush93}Rush et al. (1993) selected a 12 $\\mu$m flux-limited sample and calculated the luminosity functions for Seyfert and non-Seyfert galaxies in the sample. In a previous paper (\\markcite{paper1}Shupe et al. (1997), Paper I hereafter), we have presented the results of a 25 $\\mu$m luminosity function calculated from a large flux-limited IRAS sample containing 1456 galaxies. We continue to select a flux-limited sample and calculate the luminosity function at 12 $\\mu$m in this paper. The relation between the mid-infrared color and luminosity plays another important role in estimating various properties of galaxy evolution. It defines distinct regions in the color-flux diagram, for example, for different types of evolution and for different populations of galaxies. Such a relation is indicated by the 12-25 $\\mu$m vs. 60-100 $\\mu$m color-color relation or by the 12-25 $\\mu$m color vs. the far-infrared luminosity relation obtained from the IRAS survey (\\markcite{soi91}Soifer \\& Neugebauer 1991), and can be estimated from a large sample of galaxies selected at mid-infrared bands. Emission features near 12 $\\mu$m thought to be produced by aromatic hydrocarbon molecules have been observed in many astronomical spectra (e.g., \\markcite{gill73}Gillett et al. 1973; \\markcite{russ78}Russell et al. 1978; \\markcite{sell84}Sellgren 1984; \\markcite{roch91}Roche, Aitken, \\& Smith 1991; \\markcite{oboul96}Boulade et al. 1996; \\markcite{vig96} Vigroux et al. 1996; \\markcite{met96}Metcalfe et al. 1996; \\markcite{ces96}Cesarsky et al. 1996; \\markcite{lu97}Lu et al. 1997). These broad emission features complicate the calculations of $k$-corrections and the fluxes and luminosities at mid-infrared bands. Fortunately, ISO observations have resulted in high-quality mid-infrared spectra in various astronomical circumstances, and the on-going surveys of IRAS galaxies using ISO can provide an especially useful handle on this problem. In the next section we present a model to systematically calculate the contribution of the emission features in the mid-infrared bands of IRAS galaxies. The model is then incorporated in the following sections. Section \\ref{sec:clrlum} discusses the mid-infrared color-luminosity relation obtained from the large 25 $\\mu$m-selected sample of Paper I. The population-dependency of the relation is discussed. Then we present the calculation of the 12 $\\mu$m luminosity function in Section \\ref{sec:lumfcn}. We first discuss a selection of galaxy sample flux-limited at 12 $\\mu$m from the IRAS Faint Source Survey in Section \\ref{sec:sample}. Then the luminosity function is derived and corrected for density variations in Section \\ref{sec:pahlf}. In Section \\ref{sec:nopahlf} we discuss the effects of active galaxies and quasars on the 12 $\\mu$m luminosity function. We summarize our results in Section \\ref{sec:conclusion}. ", "conclusions": "\\label{sec:conclusion} Our main results are summarized as follows: 1. Quasars and Seyfert galaxies dominate the high luminosity regime in a 12 $\\mu$m flux-limited sample. The ultraluminous infrared galaxies are relatively rare at 12 $\\mu$m (contrast to a 25 $\\mu$m sample). 2. We have a technique for differentiating between quasars and ultraluminous infrared galaxies using their 12-25 $\\mu$m color (see Figures 5-9). Qualitatively, quasars are bluer than the luminous and ultraluminous infrared galaxies at high 25 $\\mu$m luminosities. The ultraluminous infrared galaxies also have greater far-infrared to blue luminosity ratio on average than those of the other populations. 3. A highly complete sample flux-density limited at 200 mJy at 12 $\\mu$m selected from the Faint Source Survey catalogs is used to calculate a local 12 $\\mu$m luminosity function, which is then corrected for density-variation. We are establishing a library of galaxy SEDs as a function of luminosity for more accurate $k$-corrections, and will discuss the faint source counts based on our 12 and 25 $\\mu$m luminosity functions in a forthcoming paper (Xu et al. 1997)." }, "9803/astro-ph9803024_arXiv.txt": { "abstract": "\\rxj\\ is an unidentified bright soft \\Xray\\ source which shows pulsations at a 8.39\\,s period and has a thermal spectrum. We present deep B and R band images of its \\Xray\\ localization. We find one possible counterpart in the \\Xray\\ error box, with magnitudes $B=26.6\\pm0.2$ and $R=26.9\\pm0.3$. The very high X-ray to optical flux ratio confirms that this object is an isolated neutron star. We discuss possible models and conclude that only two are consistent with the data and at the same time are able to draw from a large enough population to make finding one nearby likely. In our opinion the second criterion provides a stringent constraint but appears to have been ignored so far. The first model, suggested earlier, is that \\rxj\\ is a weakly magnetized neutron star accreting from the interstellar medium. The second is that it is a relatively young, highly magnetized neutron star, a ``magnetar'', which is kept hot by magnetic field decay. ", "introduction": "} The population of defunct radio pulsars far exceeds that of active ones. It is believed that the Galaxy has about $2\\,10^5$ radio pulsars. The neutron-star birth rate is estimated to be between one per 30 yr to one per 100 yr. Assuming a constant pulsar production rate and an age of the disk of $10^{10}$\\,yr, one infers a Galactic neutron star population of $\\sim\\!2\\,10^8$ -- three orders of magnitude larger than that of the active radio pulsar population. It is not easy to detect old neutron stars. While the nearest few intermediate-age pulsars can be identified by their cooling radiation, which peaks in the soft \\Xray/EUV band, the defunct pulsars will have become too cool to be observable. A small fraction of them, however, may be in a position to accrete matter from the interstellar medium. These will then get reheated and reappear in the \\Xray\\ sky. Quite independent of this discussion there has been a growing recognition of a population of highly magnetized neutron stars. The circumstantial evidence for this class comes from studies of soft gamma-ray repeaters (SGRs) and long-period pulsars in supernova remnants (Vasisht \\& Gotthelf \\cite{vasig:97}). Thompson \\& Duncan (\\cite{thomd:95}) have introduced the term ``magnetars'' for neutron stars with field strengths significantly larger than $10^{12}$\\,G, the typical field strength inferred for radio and \\Xray\\ pulsars. The birthrate of SGRs has been estimated to be roughly 10\\% that of the ordinary pulsars (Kulkarni \\& Frail \\cite{kulkf:93}; Kouveliotou et al.\\ \\cite{kouv&a:94}). The relevance of magnetars to the discussion at hand is as follows. Unlike the situation for ordinary neutron stars, magnetic field decay is expected to be significant in highly magnetized neutron stars. This decay could reheat the magnetar (Thompson \\& Duncan \\cite{thomd:96}), making it hotter than an ordinary neutron star, and thus brighter in soft X rays. Two of the best candidates for this general class of neutron stars have emerged from the ROSAT mission: \\rxjw\\ (Walter, Wolk, \\& Neuh\\\"auser \\cite{waltwn:96}) and \\rxj\\ (Haberl et al.\\ \\cite{habe&a:97}). Both are bright ROSAT objects with very soft \\Xray\\ spectra. Walter \\& Matthews (\\cite{waltm:97}) have provided compelling evidence for the identification of a faint blue optical counterpart of \\rxjw. In this {\\em Letter}, we present deep B and R observations of the localization of \\rxj. ", "conclusions": "" }, "9803/astro-ph9803268_arXiv.txt": { "abstract": "In the context of open inflation, we calculate the probability distribution for the density parameter $\\Omega$. A large class of two field models of open inflation do not lead to infinite open universes, but to an ensemble of inflating islands of finite size, or ``quasi-open'' universes, where the density parameter takes a range of values. Assuming we are typical observers, the models make definite predictions for the value $\\Omega$ we are most likely to observe. When compared with observations, these predictions can be used to constrain the parameters of the models. We also argue that obsevers should not be surprised to find themselves living at the time when curvature is about to dominate. ", "introduction": "Anthropic considerations have often been used in order to justify the ``naturalness'' of the values taken by certain constants of Nature \\cite{anthropic}. In these approaches, it is assumed that the ``constants'' are really random variables whose range and ``a priori'' probabilities are determined by the laws of Physics. Knowledge of these ``a priori'' probabilities is certainly useful, but not sufficient to determine the probability for an observer to measure given values of the constants. For instance, some values which are in the ``a priori'' allowed range may be incompatible with the very existence of observers, and in this case they will never be measured. The relevant question is then how to assign a weight to this selection effect. A natural framework where these ideas can be applied is inflation. There, the false-vacuum energy of the scalar field which drives the inflationary phase can thermalize in different local minima of its potential, and each local minimum may have a different set of values for the constants of Nature. Also, there may be different routes from false vacuum to a given minimum. In this case all thermalized regions will have the same low energy Physics constants, but each route will yield a hot universe with different large scale properties. Here, we shall be concerned with this possibility, where the fundamental constants (such as the gauge couplings or the cosmological constant) are fixed, but other cosmological parameters such as the density parameter or the amplitude of cosmological perturbations are random variables whose distribution is dynamically determined. In this context, the most reasonable -and predictive- version of the anthropic principle seems to be the principle of mediocrity \\cite{medi,gott}, according to which we are typical observers who shall observe what the vast majority of observers would. Thus, the measure of probability for a given set of constants is simply proportional to the total number of civilizations emerging with those values of the constants. In this paper we shall use this principle in order to calculate the probability distribution for the density parameter $\\Omega$. Standard inflationary models predict $\\Omega=1$ with ``certainty''. What this means is that these models can explain the observed homogeneity and isotropy of the universe only if the universe is flat. However, a class of ``open inflation'' models which lead to $\\Omega<1$ have received some attention in recent years \\cite{open,BGT,LM}. In these models, inflation proceeds in two steps. One starts with a scalar field $\\sigma$ trapped in a metastable minimum of its potential $V(\\sigma)$. The false vacuum energy drives an initial period of exponential expansion, and decays through quantum nucleation of highly symmetric bubbles of true vacuum. The interior of these bubbles has the geometry of an open Friedmann-Robertson-Walker universe. This accounts for the observed homogeneity and isotropy of the universe. In order to solve the flatness problem a second stage of slow roll inflation inside the bubble is necessary. In models with a single scalar field $\\sigma$, all bubbles have the same value of $\\Omega$ which is determined by the number of e-foldings in the second period of inflation. The potential $V(\\sigma)$ in such models is assumed to have a rather special form, with a sharp barrier next to a flat slow-roll region, which requires a substantial amount of fine-tuning. Additional tunning is needed to arrange the desired value of $\\Omega$. A more natural class of models includes two fields, $\\sigma$ and $\\phi$, with $\\sigma$ doing the tunneling and $\\phi$ the slow roll \\cite{LM}. The simplest example is \\begin{equation} \\label{coupled} V(\\sigma,\\phi)=V_t(\\sigma) + {g \\over 2} \\sigma^2 \\phi^2, \\end{equation} where $V_0(\\sigma)$ has a metastable false vacuum at $\\sigma=0$. After $\\sigma$ tunnels to its true minimum $\\sigma=v$, the field $\\phi$ would drive a second period of slow roll inflation inside the bubble. Depending on the value of $\\phi$ at the time of nucleation, the number of e-foldings of the second stage of inflation would be different. Initially, it was believed \\cite{LM} that models such as (\\ref{coupled}) would yield an ensemble of infinite open universes, one inside each nucleated bubble, and each one with a different value of the density parameter. However, it has been recently realized \\cite{GGM} that this picture is oversimplified. The two field models which allow for variable $\\Omega$ do not actually lead to infinite open universes, but to an ensemble of inflating islands of finite size inside of each bubble. These islands are called quasi-open universes. Within each island, the number of e-foldings of inflation decreases as we move from the center to the edges. Also, each island is characterized by a different number of e-foldings in its central region. As a result, even within the same bubble, different observers will measure a range of values of the density parameter. The picture of the large scale structure of the universe in these models is rather simple, because all bubbles have the same statistical properties. We shall see that the quasiopen nature of inflation is of crucial importance for the calculation of the probability distribution for the density parameter. In models of quasiopen inflation, such as (\\ref{coupled}), $\\Omega$ takes different values in different parts of the universe, while the other constants of Nature and cosmological parameters remain fixed. More general models can be constructed where other parameters can change as well, and in Section VII we give an example of a model with a variable amplitude of density fluctuations. However, our main focus in this paper is on the models in which only $\\Omega$ is allowed to vary. In order to apply the principle of mediocrity to our models, we will have to compare the number of civilizations in parts of the universe with different values of $\\Omega$. Of course, we cannot calculate the number of civilizations. However, since the value of $\\Omega$ does not affect the physical precesses involved in the evolution of life, this number must be proportional to the number of habitable stars or, as a rough approximation, to the number of galaxies. Hence, we shall set the probability for us to observe a certain value of $\\Omega$ to be proportional to the number of galaxies formed in parts of the universe where $\\Omega$ takes the specified value. The principle of mediocrity was applied to calculate the probability distribution for $\\Omega$ in an earlier paper \\cite{VW}, which assumed the old picture of homogeneous open universes inside bubbles. A serious difficulty encountered in that calculation was that open universes inside the bubbles have infinite volume and contain an infinite number of galaxies. Thus, to find the relative probability for different values of $\\Omega$, one had to compare infinities, which is an inherently ambiguous task. This problem was addressed in \\cite{VW} by introducing a cutoff and counting only galaxies formed prior to the cutoff. Although the cutoff procedure employed in \\cite{VW} has some nice properties, it is not unique, and the resulting probability distribution is sensitive to the choice of cutoff \\cite{linde}. This cutoff dependence, which also appears in other models of eternal inflation \\cite{linde,vireg}, has lead some authors to doubt that a meaningful definition of probabilities in such models is even in principle possible \\cite{linde,gbl}. However, this pessimistic conclusion may have been premature. According to the quasiopen picture, $\\Omega$ takes all its possible values within each bubble. Since all bubbles are statistically equivalent, it is sufficient to consider a single bubble. Moreover, we can restrict ourselves to a finite (but very large) comoving volume within that bubble, provided that its size is much greater than the characteristic scale of variation of $\\Omega$. Thus, we no longer need to compare infinities, and the problem becomes well defined. The possibility of unambiguous calculation of probabilities in the quasiopen model was our main motivation for revising the analysis of Ref.\\cite{VW}. Also, we shall give a more careful treatment of the astrophysical aspects of the problem which were discussed rather sketchily in \\cite{VW}. The paper is organized as follows. In Section II we review the main features of quasi-open inflation. In Section III we introduce the probability distribution for $\\Omega$. A basic ingredient in this distribution will be the anthropic factor $\\nu(\\Omega)$, which gives the number of civilizations that develop per unit thermalized volume in a region characterized by a certain value of $\\Omega$. In Section IV we evaluate $\\nu(\\Omega)$ and calculate the probability distribution for $\\Omega$ in the model (\\ref{coupled}). In Section V we extend our results to more general models with arbitrary slow roll potentials for the field $\\phi$. In Section VI we discuss observational constraints on quasiopen models due to CMB anisotropies and how these constraints restrict the class of models that give a probability distribution peaked at a non-trivial value of $\\Omega$. In Section VII we comment on the ``cosmic age coincidence'', that is, on whether it would be surprising to find ourselves living at the time when the curvature of the universe starts dominating. In Section VIII we summarize our conclusions. Some side issues and technical details are discussed in the appendices. ", "conclusions": "We have calculated the probability distribution for the density parameter in models of open inflation with variable $\\Omega$. This probability is basically the product of three factors: the ``tunneling'' factor, which is related to the microphysics of bubble nucleation and subsequent expansion; the volume factor, related to the amount of slow roll inflation undergone in different regions of the universe; and the ``anthropic factor'', which determines the number of galaxies that will develop per unit thermalized volume. It is interesting that the expression for the probability (\\ref{distributiony}) depends on the underlying particle physics model through a single dimensionless parameter $\\mu$, defined in Eq.(\\ref{defmu}). Taking the minimum of the slow roll potential to be at $\\phi=0$, the tunneling factor tends to suppress large initial values of $\\phi$, favouring low values of $\\Omega$. However, only those regions for which $\\phi$ is large enough will inflate. Hence, there will be a competition between volume enhancement and ``tunneling'' suppression. The most interesting situation occurs when the tunneling suppression dominates over the volume factor. In this case, the product of both would peak at $\\Omega=0$, and the anthropic factor $\\nu(\\Omega)$ becomes essential in determining the probability distribution. In an open universe, cosmological perturbations stop growing when the universe becomes curvature dominated, and for low values of $\\Omega$ structure formation is suppressed. The effect of the anthropic factor is, therefore, to shift the peak of the distribution from $\\Omega=0$ to a nonzero value of $\\Omega$. As a first approximation \\cite{VW,MSW}, we have taken $\\nu(\\Omega)$ to be proportional to the fraction of matter that clusters on the galactic mass scale in the entire history of a certain region. We have found that the peak of the distribution is given by the condition \\begin{equation} \\kappa \\left({1-\\Omega \\over \\Omega}\\right)_{peak} \\approx \\left({3\\over 2}\\mu-{5\\over 4}\\right)^{1/2}, \\label{mon} \\end{equation} where the coefficient $\\kappa \\sim 10^{-1}$ is defined in (\\ref{kappa}). For models with $\\mu \\sim 1$ (which can be easily constructed), the probablility distribution for the density parameter ${\\cal P}(\\Omega)$ can peak at values of $\\Omega$ such that $x=(1-\\Omega)/\\Omega\\sim 1$ (See Fig. 1). The peaks are not too sharp, with amplitude $\\Delta y \\approx 1/2$, or $\\Delta x \\approx 5$, so a range of values of $\\Omega$ would be measured by typical observers. The analysis we presented here demonstrates that, given a particle physics model, the probability distribution for $\\Omega$ can be unambiguously calculated from first principles. We can also invert this approach and use our results to exclude particle physics models which give the peak of the distribution at unacceptably low values of $\\Omega$. This gives the constraint $\\mu\\lesssim 3$. An independent constraint on the model parameters can be obtained from CMB observations. If the observed CMB anisotropies are to be explained within the same two-field model of open inflation, without adding any extra fields, then we have shown in Section IV that the corresponding constraint (if the observed value of $\\Omega$ lies in the range $.1$ to $.7$) is $\\mu \\gtrsim 10^{6} \\epsilon^2$, where $\\epsilon$ is the slow roll parameter defined in (\\ref{epsilon}). Combinig both constraints, we obtain a bound on the slow roll parameter $$ \\epsilon\\lesssim 10^{-3}. $$ This bound is somewhat restrictive. For instance, for the simple free field model (\\ref{coupled}), the slow roll parameter is of order $10^{-2}$, and so this model would contradict observations. It is easy, however, to generalize the slow roll potential in order to make $\\epsilon$ sufficiently small. If one allows some other source for CMB fluctuations (e.g., topological defects), then the CMB constraint is much less restrictive, and simple models of the form (\\ref{coupled}) are still viable. We have advanced anthropic arguments towards explaining the ``cosmic age coincidence'', that is, whether it would be surprising to find that we live at the time when the curvature is about to dominate. We have argued that this is not unexpected. We have also discussed a three-field model in which the amplitude of density fluctuations $Q$ becomes a random variable. We have outlined an argument explaining the observed value $Q\\sim 10^{-5}$ in the framework of this model. While this work was being completed, Hawking and Turok \\cite{HT98}, have suggested the possibility of creation of an open universe from nothing (see also \\cite{everybody}). The validity of the instantons describing this process \\cite{alex}, and also their ability to successfully reproduce a sufficiently homogeneous universe, is still a matter of debate and needs further investigation. Clearly, the analysis presented in this paper can be easily adapted to this new framework." }, "9803/astro-ph9803118_arXiv.txt": { "abstract": "We use the two-point correlation function to calculate the clustering properties of the recently completed SSRS2 survey, which probes two well separated regions of the sky, allowing one to evaluate the sensitivity of sample-to-sample variations. Taking advantage of the large number of galaxies in the combined sample, we also investigate the dependence of clustering on the internal properties of galaxies. The redshift space correlation function for the combined magnitude-limited sample of the SSRS2 is given by $\\xi(s)=(s/5.85$ \\h1 Mpc)$^{-1.60}$ for separations between 2 $\\leq s \\leq$ 11 \\h1 Mpc, while our best estimate for the real space correlation function is $\\xi (r) = (r/5.36$ \\h1 Mpc)$^{-1.86}$. Both are comparable to previous measurements using surveys of optical galaxies over much larger and independent volumes. By comparing the correlation function calculated in redshift and real space we find that the redshift distortion on intermediate scales is small. This result implies that the observed redshift-space distribution of galaxies is close to that in real space, and that $\\beta = \\Omega^{0.6}/b < 1$, where $\\Omega$ is the cosmological density parameter and $b$ is the linear biasing factor for optical galaxies. We have used the SSRS2 sample to study the dependence of $\\xi$ on the internal properties of galaxies such as luminosity, morphology and color. We confirm earlier results that luminous galaxies ($L>L^*$) are more clustered than sub-$L^*$ galaxies and that the luminosity segregation is scale-independent. We also find that early types are more clustered than late types. However, in the absence of rich clusters, the relative bias between early and late types in real space, $b_{E+S0}$/$b_S$ $\\sim$ 1.2, is not as strong as previously estimated. Furthermore, both morphologies present a luminosity-dependent bias, with the early types showing a slightly stronger dependence on the luminosity. We also find that red galaxies are significantly more clustered than blue ones, with a mean relative bias of $b_R/b_B$ $\\sim$ 1.4, stronger than that seen for morphology. Finally, by comparing our results with the measurements obtained from the infrared-selected galaxies we determine that the relative bias between optical and \\iras galaxies in real space is $b_o/b_I$ $\\sim$ 1.4. ", "introduction": "\\subsection{Method} The two-point correlation function $\\xi (r)$ can be computed from the data using the estimator suggested by Hamilton (1993): \\begin{equation} \\xi(r) = {DD(r) RR(r) \\over [DR(r)]^2} - 1, \\end{equation} where $DD(r)$, $RR(r)$ and $DR(r)$ are the number of data--data, random--random and data--random pairs, with separations in the interval between $r$ and $r + dr$. The random catalog is generated using the same selection criteria as the galaxy sample. This estimator has the advantage that it is not too sensitive to uncertainties in the mean density, which is only a second order effect. The counts $DD(r)$, $DR(r)$ and $RR(r)$ can be generalized to include a weight $w$ which is particularly important to correct for selection effects at large distances in magnitude-limited samples: \\begin{eqnarray} DD(r) = &{\\displaystyle{ \\sum_i^{N_{\\scriptscriptstyle{gal}}} \\sum_j^{N_{\\scriptscriptstyle{gal}}}} } & w(s_j, r) w(s_i, r), \\\\ & \\scriptscriptstyle {r- \\Delta r \\leq |s_i-s_j| \\leq r+ \\Delta r} \\nonumber \\end{eqnarray} where $i$ sums over all objects in the sample and the sum over $j$ includes all particles at a distance $s$ from the origin, which in this work is taken as the centroid of the Local Group, and $r = | {\\bf{s}}_i -{\\bf{s}}_j|$ is the separation of the pair $(i,j)$. The galaxy-random pairs $DR(r)$ and random-random pairs $RR(r)$ are similarly weighted. The most common weighting schemes are: equally weighted pairs $w(s_i, r)=1$; equally--weighted volumes where $w(s_i, r)=1/\\phi(s_i)$ and the minimum variance weighting given by \\begin{equation} w (s_i, r) = {1 \\over 1 +4\\pi \\bar n J_3(r) \\phi(s_i)}, \\quad J_3 (r) = \\int_0^r dr'r'^2 \\xi(r'), \\end{equation} where $\\phi (s_i)$ is the selection function at distance $s_i$ from the origin and $J_3$ is the mean number of excess galaxies out to a distance $r$ around each galaxy. Even though in the last scheme the weights depend on the unknown correlation function, in practice, it is not very sensitive to the exact form of $\\xi(r)$ (\\eg Loveday \\etal 1992; Marzke, Huchra \\& Geller 1994). In this work we adopt the minimum-variance weighting and take $J_3 (r$ = 30 \\h1 Mpc) $\\sim 1100$, obtained by using the real-space correlation function of Davis \\& Peebles (1983). The mean densities were calculated using the estimator \\begin{equation} \\bar n = \\sum_{i=1}^{N_{gal}}w_i/\\int_{s_{min}}^{s_{max}} dV\\phi(s)w(s) \\end{equation} where again $\\phi(s)$ is the selection function, derived from the luminosity function and $w(s)$ is the weight (\\eg Davis \\& Huchra 1982). The errors for the redshift space correlation function (\\xis) as well as for the real-space correlation discussed below, were calculated by means of bootstrap resampling (Ling, Frenk \\& Barrow 1986). For the volume-limited samples the total of bootstraps was 50, while for magnitude-limited samples 25 resamplings were calculated. As shown by Fisher \\etal (1994), bootstraping tends to overestimate the true errors, so that the estimate of the latter will in general be rather conservative. \\subsection{Real space} In order to estimate real space correlation functions, we follow Davis \\& Peebles (1983). For any two galaxies with redshifts {\\bf s$_1$} and {\\bf s$_2$}, we define the separation in redshift space, and the separation perpendicular to the line of sight respectively as \\begin{equation} {\\bf {s = s_1 - s_2}}, \\quad {\\bf {l}} = {1 \\over 2} \\bf{(s_1 + s_2)}, \\end{equation} in the small angle approximation. From these parameters one can derive $\\pi$, the separation between two galaxies parallel to the line of sight and $r_p$, the separation perpendicular to the line of sight using: \\begin{equation} \\pi = {{\\bf s.l} \\over |l|}, \\quad r_p = \\sqrt{|{\\bf{s}}|^2 -\\pi^2}. \\end{equation} These are then used to compute the statistic $\\xi(r_p,\\pi)$ estimated from the pair-counts as \\begin{equation} 1 + \\xi(r_p,\\pi) = {DD(r_p,\\pi) RR(r_p,\\pi) \\over [DR(r_p,\\pi)]^2}. \\end{equation} From \\xip we define the projected function: \\begin{equation} \\omega(r_p) = 2 \\int_0^\\infty d\\pi \\quad \\xi (r_p,\\pi), \\end{equation} which is related to the real space correlation function through \\begin{equation} \\omega(r_p) = 2\\int_0^\\infty dy \\quad \\xi[(r_p^2 + y^2)^{1/2}]. \\end{equation} The inverse is the Abel integral: \\begin{equation} \\xi(r) = -{1 \\over \\pi} \\int_r^\\infty dr_p {\\omega^\\prime(r_p) \\over (r_p^2-r^2)^{1/2}, } \\end{equation} where $\\omega^\\prime(r_p)$ is the first derivative of $\\omega (r_p)$. If the real space correlation function is a power-law, the integral for $\\omega(r_p)$ can be performed analytically to give \\begin{equation} \\omega(r_p) = r_p ({r_o \\over r_p})^{\\gamma} { \\Gamma({1 \\over 2}) \\Gamma({\\gamma -1 \\over 2}) \\over \\Gamma({\\gamma \\over 2})}. \\end{equation} \\subsection {Biasing} The variance of galaxy counts measures the clustering amplitude at intermediate scales. It is also a useful quantity to compare models and data. The variance in the counts is defined as \\begin{equation} \\langle (N -nV)^2 \\rangle = nV +n^2V^2\\sigma^2, \\end{equation} where nV is the mean number of galaxies in the volume V and $ n^2V^2\\sigma^2$ is the mean number of galaxies in excess of random inside a sphere of volume V. It is related to the moment of the correlation function (Peebles 1980) \\begin{equation} \\sigma^2 = {1 \\over V^2} \\int_VdV_1dV_2 \\xi(|r_1-r_2|), \\end{equation} which can be calculated numerically. For a power law correlation function $\\xi(r) = (r/r_o)^\\gamma$, and a spherical volume of radius R we get \\begin{equation} \\sigma^2(R) = 72(r_o/R)^\\gamma/\\lbrack 2^\\gamma(3-\\gamma)(4-\\gamma)(6-\\gamma)\\rbrack. \\end{equation} This is the expression we have used to compute \\sig8 , and which is often used to normalize theoretical models. The relative bias between two different samples at a given separation $s$ may be estimated through (\\eg Benoist et al. 1996) : \\begin{equation} \\frac {b} {b_*}(s) = \\sqrt{\\frac{\\xi(s)}{\\xi_*(s)}} = \\sqrt{\\frac{J_3(s)}{J_3*(s)}}, \\end{equation} where the starred symbols denote a sample taken as a fiducial. The relative bias of the clustering may also be estimated through \\begin{equation} \\frac {b} {b_*}(s) = \\sqrt{\\frac{\\sigma^2(s)}{\\sigma_*^2(s)}}, \\end{equation} where $\\sigma^2$ is the variance of counts in cells described above. These are the expressions used in this work to calculate the relative bias between galaxies of different luminosities relative to $L^*$ galaxies, as well as for different morphological types and colors. \\section {Magnitude-limited Samples} In order to estimate the effects due to the finite volume we are probing and to estimate the importance of cosmic variance, we compare the clustering properties of the individual SSRS2 south and north samples as well as the combined sample, with previous estimates of $\\xi$. In this analysis, we have computed \\xis taking into account all galaxies brighter than $M=-13$ in the velocity range $500 < v < 12,000$ \\kms. The correlation function was computed using the minimum-variance weighting discussed in Section 3 and a random background catalog of 10,000 points for the individual samples and 20,000 points for the combined sample. In the calculation of the selection function we have used the Schechter parameters determined for the entire SSRS2 survey by da Costa et al. (1997), which are $M^*$ = -19.55 and $\\alpha$ = -1.15. These values are virtually identical to those measured by da Costa et al. (1994) for the SSRS2 south. We tested whether our results are affected by the presence of clusters of galaxies. For this we used a list of galaxy clusters with richness R $\\geq$ 1 (J. Huchra, private communication). All galaxies whose positions were within one Abell radius of the central position of cluster, and that had radial velocities within 500 \\kms of the cluster's mean radial velocity were culled from the sample. We find that the correlation parameters are virtually unchanged for the vast majority of the samples, and when there are changes, these are within the quoted errors of the complete sample. Therefore, we will not consider the removal of galaxies in clusters in this work. The redshift space correlation function, \\xis, for the SSRS2 samples is shown in Fig. 1, where we plot the correlation function out to separations of 30 \\h1 Mpc. For the sake of clarity, in the figure we only show error bars calculated for the combined sample. One can see that beyond $\\sim$ 15 \\h1 Mpc, the errors become progressively larger, and sometimes the sample-to-sample variations are larger than the estimated errors. In general, \\xis is adequately described by a power-law on small scales. For most cases in this paper, the power-law fits were calculated in the interval $2 < s < 11$ \\h1 Mpc. The upper-limit was chosen because there is a suggestion of an abrupt change of slope in \\xis on scales $s$ $\\lsim$ 12 \\h1 Mpc. The lower-limit was chosen to minimize the effects on \\xis due to peculiar motions of galaxies in virialized systems. The best power-law fits obtained for each sub-sample of the SSRS2 are represented as lines in Fig. 1, as explained in the caption. The correlation parameters derived from the fits are presented in Table 1, where we list: the sample identification (column 1); the correlation length (column 2) and slope $\\gamma_s$ (column 3) obtained from the power-law fits; and in column (4) the rms variance in galaxy counts within spheres 8 \\h1 Mpc in radius, followed in columns (5) through (7) by the same parameters determined for real space, which will be discussed below. An inspection of both Table 1 and Fig. 1 shows that the redshift correlation functions for SSRS2 sub-samples are very similar on small scales ($s < 10$ \\h1 Mpc). This also demonstrates that the sampling variations are consistent with the error estimates, at least in the range of separations for which the fits are calculated. In Fig. 2 we compare \\xis measured in this work for the combined SSRS2 with the \\xis measured in other surveys - the sparsely sampled Stromlo-APM survey (Loveday et al. 1995), the Las Campanas Redshift Survey (LCRS) (Tucker \\etal 1997), and that measured by Fisher et al. (1994) for the 1.2 Jy \\iras survey. The fit parameters calculated in these papers, as well as by other workers can be found in Table 2. Despite small differences in amplitude, the shapes of the three optical surveys are remarkably similar. It is important to note that the volumes of the Stromlo-APM ( 2.5 $\\times$ 10$^6$ $h^{-3}$ Mpc$^3$) and the LCRS ( 2.6 $\\times$ 10$^6$ $h^{-3}$ Mpc$^3$) are about 5 times larger than that of the SSRS2 (5.2 $\\times$ 10$^5$ $h^{-3}$ Mpc$^3$), and probe different regions of space, and thus independent structures. The lower amplitude of the \\iras survey compared to the optical samples reflects the relative bias that exists between optically and infrared-selected galaxies, which will be further discussed in Section 6 below. In Fig. 3, we compare our power-law fit parameters with equivalent measurements by other authors (see Table 2). In general, there is a good agreement between our values for the redshift space parameters and those obtained from other optical surveys, specially the Stromlo-APM and LCRS. The effect of redshift distortions on the observed redshift correlation function has also been estimated for the SSRS2. These distortions are caused by the peculiar velocities of galaxies, which on large scales, are due to the infall of galaxies from low-density regions into high-density regions, while on small scales the correlations are smeared out by virial motions of galaxies in groups and clusters (\\eg Kaiser 1987). As described in Section 3, these effects may be accounted for by calculating the correlation function as a function of the separations parallel and perpendicular to the line of sight, which can then be used to define the \\wrp estimator, which is unaffected by redshift distortions. However, one should bear in mind that in general the calculation of the real space correlation function is much more susceptible to noise than that calculated in redshift space. From the correlation functions \\xip computed using the minimum-variance weighting scheme, we have obtained \\wrp. From power-law fits, in the interval 2 $< r_p <$10 \\h1 Mpc, we have derived the correlation parameters listed in Table 1. By comparing the real space fit parameters obtained in this work (Table 1) with previous measures (columns 4 and 5 in Table 2), we find a good agreement with the real space measurements of Davis \\& Peebles (1983), Loveday et al. (1995) and Marzke et al. (1995). The fit to the real space correlation function for the combined sample is compared in Fig. 4 with the redshift space correlation function. One can see that at intermediate separations the redshift \\xis is amplified relative to the real space correlation \\xir. The small amplification suggests that the observed redshift distribution is close to the real space distribution. At separations of $\\sim$ 10 \\h1 Mpc, the ratio between the real and redshift space correlations is $\\sim$ 1.5. In the linear regime, peculiar motions on large scales cause \\xir to be amplified by a factor $\\sim$ $1 + { 1 \\over 2 } {\\beta} + { 1 \\over 5 } {\\beta^2}$ where $\\beta = {\\Omega^{0.6} \\over b}$ and $b$ is the linear biasing factor (Kaiser 1987). Therefore, a rough estimate for $\\beta$ is $ \\sim 0.6$, on scales of the order of 10 \\h1 Mpc, consistent with that determined by Loveday \\etal (1996). \\section {The Clustering Dependence on the Internal Properties of Galaxies} \\subsection {Luminosity} In this work we use the combined SSRS2, as well as the SSRS2 north and south sub-samples to further explore the clustering dependence on luminosity, as was carried out by Benoist et al. (1996), but who only used the SSRS2 south. Probing independent structures in different volumes we can estimate the impact of cosmic variance. It should also be noted that the absolute magnitude limits considered in this section differ slightly from those of Benoist et al. (1996), and were chosen to compare our results with the volume-limited samples of Fisher \\etal (1994), which will be discussed in Section 6 below. The volume-limited samples considered in this section only contain galaxies bright enough that would allow them to be included in the sample when placed at the cutoff distance. We defined samples limited at radial distances of 60, 80, 100 and 120 \\h1 Mpc. The absolute magnitude limits corresponding to these distances are -18.39, -19.01 (both $L < L^*$) , -19.50 ($\\sim L^*$) and -19.89 ($L > L^*$), respectively. For all galaxies in these samples, the weighting function is $w(r)=1$ and the volume densities are simply the total number of galaxies divided by the corresponding volume. The correlation functions obtained for the volume-limited sub-samples at the different depths are shown in Fig. 5 for $s \\leq 20$ \\h1 Mpc, where the different symbols represent different volume limits. For reasons of clarity, we only present error bars for the samples volume-limited at 60 \\h1 and 120 \\h1 Mpc. The meaning of these symbols, as well as the indication of the parent sample (SSRS2 south, north or combined) are shown in each panel. The power-law fits are represented by lines in the figure and the parameters are summarized in Table 3 where we list: in column (1) the sample; in column (2) the depth R; in column (3) the number of galaxies N$_g$; in column (4) the mean density; in columns (5) and (6) the power-law fit parameters and formal errors; and in column (7) $\\sigma_8$, the rms fluctuation of the number of galaxies in a sphere of radius 8\\h1 Mpc. The interval used in the fits is $ 2 < s < 11$ \\h1 Mpc, the same as that adopted in the previous section. An inspection of Table 3 and Fig. 5 shows that the amplitude of \\xis tends to increase with the sample depth, the variation being somewhat larger in the northern and combined samples. We point out that \\xis for the SSRS2 north (panel b), is noisier because of the smaller number of galaxies, in most cases about half of those in the southern sample. The correlation length ($s_0$) ranges from 3.8 \\h1 Mpc to 6.8 \\h1 Mpc. However, the slope varies considerably from sample to sample, though with a tendency of becoming steeper as the depth increases. In order to evaluate the cosmic variance, we show in Fig. 6 \\xis for each of the volume-limited sub-samples, but now plotting the results for the southern, northern and combined samples in each panel. For the samples in smaller volumes, the differences between the northern and southern samples are larger than the estimated error calculated for the combined sample, and probably reflect the amplitude of the sample to sample variation, with the north being systematically lower. For the larger volumes the samples present similar behavior, and the variations are generally consistent with the estimated errors. To remove the effects of distortions due to motions, which may affect our estimates of the strength of clustering and the relative bias between different samples, we have also computed the real-space correlation function for the volume-limited samples. As above, we have computed \\xip for the sub-samples volume-limited at R = 60, 80, 100 and 120 \\h1 Mpc in each galactic hemisphere and for the combined sample. The resulting real space correlation parameters are listed in Table 4. In Fig. 7 we compare \\xis measured for each volume limit, denoted by open symbols, with the real-space correlation fits described above, represented as a solid line. For the sake of clarity, we only show the fits we measure for the combined sample, as this will be the one less affected by noise. The smearing due to motions in virialized systems for $r < 3$ \\h1 Mpc is quite noticeable for all samples, while the effect of peculiar motions is only obvious for the smaller volumes, little evidence being seen in the samples in larger volumes. The dependence of clustering in redshift--space (as measured by $\\sigma_8$) with luminosity (as measured by the limiting absolute magnitude of each sample) is shown in Fig. 8(a), where we use as fiducial magnitude the value of $M^*$=-19.55 (see Section 4). The figure shows an overall behavior consistent with that found by Benoist et al. (1996) and which is detected in all samples, further demonstrating that this effect is unlikely to be spurious. This result supports their finding that there is a dependence of clustering on luminosity, as measured in redshift space. To further investigate its reality, we have computed \\xir in real space for the same volume limited samples. The results are shown in Fig. 8(b). Here again it is immediately apparent that the clustering amplitude increases with luminosity in the same way as seen in redshift space. On the whole, these results, using a larger sample, confirm in real space the conclusions of Benoist \\etal (1996). In Fig. 9 we present the relative bias with scale calculated using equation (15), where we compare the the correlation function for the volume-limited samples at 60, 80 and 120 \\h1 Mpc relative to the 100 \\h1 Mpc sample. From the figure one may see that there are only minor differences between the smaller volumes. In the case of the sample volume-limited at 120 \\h1 Mpc, the relative bias is fairly constant over the range of scales we consider at $\\sim$ 1.5. This suggests that the luminosity bias is scale-independent, and that it starts to become important only for galaxies brighter than $\\sim$ $L^*$. \\subsection{Morphology} Since all galaxies in the SSRS2 have morphological classifications we can also analyze the clustering dependence on morphology. With this aim, we have calculated \\xis and $\\xi(r)$ for the SSRS2 for different morphological types, dividing galaxies into broad morphological bins - early types comprising E, S0 and S0-a, and late types containing Sa galaxies and later. In contrast to the results of the Stromlo-APM, the luminosity function parameters used in the selection function for both samples are quite similar to those measured for the SSRS2 as a whole (Marzke et al. 1997). Furthermore, since sample-to-sample variations are within our estimated errors both for the magnitude and volume-limited samples, as shown in Section 4, in the analysis below we only consider the combined sample to improve the statistics. The resulting correlation functions for early and late type galaxies are presented in Figure 10 panel (a) in redshift space and (b) in real space, while the fit parameters are presented in Table 5. For the late type galaxies we find that the correlation function is adequately described by ($s_0$ = 5.4\\mm0.2\\h1 Mpc; $\\gamma_s$=1.48\\mm0.09), while for early types we find $s_0$ = 6.5\\mm0.2 \\h1 Mpc; $\\gamma_s$=1.86\\mm0.11. Our values for early type galaxies are close to those of Santiago \\& da Costa for the diameter-limited SSRS ($s_0$ = 6.0\\mm1.5 \\h1 Mpc, $\\gamma_s$=1.69) and Hermit \\etal (1996) for the ORS, who measure $s_0$=6.7 and $\\gamma_s$=1.52. Our value for late types is somewhat larger than that measured by Santiago \\& da Costa (1990), while a proper comparison with Hermit \\etal (1996) cannot be made, because we have not subdivided spirals into earlier (Sa/Sb) and later (Sc/Sd) types as they did. A comparison between the fit parameters obtained from available redshift space correlation functions, is shown in Fig. 11, where open symbols represent fits for late type galaxies and solid symbols represent early types. Although all works agree that early types are more clustered than late types, as indicated by the larger correlation length, the scatter is large with the Stromlo-APM results yielding very extreme results. This, in turn, implies large uncertainties in the measurement of the relative bias between the two populations. Based on the redshift space information, we estimate the relative bias between morphological types as 1.25. However, for a proper estimate of the dependence of the correlation properties on morphology it is important to take into account the fact that redshift distortions may affect early and late type galaxies in different ways. Therefore a more meaningful comparison must be carried out in real space. The values we measure for the correlation length in real space for early types ($r_0$=6.0\\mm0.4; $\\gamma_r$=1.91\\mm0.18) show a very good agreement with those of Loveday \\etal (1995), ($r_0$=5.9\\mm0.7; $\\gamma_r$=1.85\\mm0.13). For late types we find ($r_0$=5.3\\mm0.3; $\\gamma_r$=1.89\\mm0.15), which is somewhat larger than those measured by the same authors ($r_0$=4.4\\mm0.1; $\\gamma_r$=1.64\\mm0.05). Our value of the correlation length is significantly smaller than that measured by Guzzo et al. (1997) ($r_0$=8.4\\mm0.8; $\\gamma_r$=2.05\\mm0.09) for early types. We should note that their sample is volume-limited at $M < -19.5$, whereas we consider galaxies down to $M = -13$. In order to compare with these authors we consider a volume-limited sub-sample of SSRS2 galaxies with $M \\leq$ -19.5, which corresponds to maximum distance of 100 \\h1 Mpc. Using this sample, for early types we measure ($r_0$=5.7\\mm0.8; $\\gamma_r$=2.09\\mm0.49) while for late types we find ($r_0$=5.0\\mm0.5; $\\gamma_r$=2.01\\mm0.28). For both early as well as late types, there are still discrepancies relative to the results of Guzzo et al. (1997), which could reflect the paucity of rich clusters in our sample. By using the variance, we estimate that the relative bias between the different morphologies is $b_{E+S0}/b_S$ = 1.18$\\pm$0.15 in a sample where clusters are not important. This value is smaller than the determination derived from the real-space correlations of Loveday et al. (1995) $b_{E+S0}/b_S$ = 1.33 and Guzzo et al. (1997) $b_{E+S0}/b_S$ = 1.68. From these results we may conclude that the relative bias between the two populations range from roughly 1.2 to 1.7, depending on the cluster abundance in the sample, with the former value representing a lower limit. We have also calculated the correlation function for galaxies discriminated by morphological types for volume-limited samples using the same absolute magnitude limits as in Section 5.1. This calculation was carried out both in redshift, as well as real space, and the results are presented in Tables 6 and 7 respectively. In redshift space there is a trend of the correlation function amplitude increasing with luminosity for both morphological classes. The magnitude of this variation is larger for early types than for late types, although the errors are large. The same trend may be inferred from the analysis in real space, as shown in Figure 12(a), where we compare the $\\sigma_8$ values obtained for the different sub-samples. Here it may be clearly seen that there is a trend of $\\sigma_8$ increasing with luminosity, suggesting that the morphological and luminosity segregations are two separate effects. Using equation (15) we can also examine how the relative bias varies as a function of scale. This is shown in Figure 12(b), using the real space correlation functions. In contrast to the luminosity bias we find that the morphological bias presents a small decrease from $\\sim$ 1.4 on small scales to $\\sim$ 1.0 on larger scales ($\\sim$ 8 \\h1 Mpc). Although the latter value is slightly smaller than that estimated through the $\\sigma_8$ values ($b_{E+S0}/b_S$ = 1.18 $\\pm$ 0.15), it is still within the estimated error. A similar behavior of the morphological bias changing with scale, was found by Hermit et al. (1996) but using the redshift space correlation function of the ORS, which may not be as meaningful, because of possible biases introduced by virial motions. Taken together, the above results are consistent with the interpretation that luminosity segregation could be a primordial effect, while the morphological segregation could be enhanced by environmental effects (e.g. Loveday \\etal 1995). \\subsection{Colors} Another internal characteristic available in the present catalog is color. Although morphology and colors are correlated the scatter is large, and galaxies of a given type exhibit a broad range of colors, indicating different star-formation histories. On the other hand, colors are easily measured and are an objective criterion, in particular for samples of distant galaxies, whereas the morphological classification is somewhat subjective and becomes increasingly difficult to carry out as the apparent sizes of galaxies get smaller. A further evidence that morphology and colors have somewhat different distributions comes from the calculation of the luminosity function, which presents significantly different shapes for blue and red galaxies (Marzke \\& da Costa 1997), while the luminosity function calculated by separating galaxies between early and late types presents similar Schechter parameters (Marzke et al. 1997). The few works calculating the correlation properties of galaxies divided by colors present rather conflicting results for the deep samples. Works by Infante \\& Pritchet (1993) and Landy, Szalay \\& Koo (1996) using the angular correlation function show that the correlation of redder galaxies is significantly stronger than for bluer galaxies, except for the very bluest ones (Landy et al. 1996). Carlberg et al. (1996) analyzing a redshift survey of K-band selected galaxies, find that for $0.3 \\leq z \\leq 0.9$ red galaxies are more correlated than blue galaxies by a factor of five. These results differ from those of Le F\\`evre et al. (1996) who find that at $z \\geq$ 0.5 blue and red galaxies have the same correlation properties, while for 0.2 $\\leq z \\leq$ 0.5 blue galaxies are less correlated than red ones. For nearby galaxies, Tucker et al. (1996) have calculated the correlation function and showed that at small scales ($s \\leq 10$ \\h1 Mpc) red galaxies ($[b_J - R]_0 > 1.25$) cluster more strongly than blue ($[b_J - R]_0 < 1.05$) ones, while for larger scales no evidence of color segregation is seen. In order to make an independent estimation of the dependence of \\xis on colors, we use the the $m_B$ = 14.5 sample described in Section 2, which contains galaxies in both galactic hemispheres. As mentioned in Section 2, this bright limit was used because of incompleteness in colors, as we are restricted to galaxies with measurements in the Lauberts \\& Valentijn (1989) catalog. In this work we adopted the restframe color cutoff as $(B_T-R_T)_0$ = 1.3 which is roughly the color of an Sbc galaxy, and was the criterion adopted by Marzke \\& da Costa (1997) in the determination of the luminosity function by colors. This value is close to the median value of $B_T-R_T$ in our sample which is $B_T-R_T$=1.2. The conversion of observed into restframe colors used the no-evolution models calculated by Bruzual \\& Charlot (1993), where we assume that the B and R measures in the Lauberts \\& Valentijn (1989) catalog are on the same system of $b_J$ and $r_F$ used by Bruzual \\& Charlot (1993). To calculate \\xis we used the following Schechter function parameters; for blue galaxies ($B_T-R_T \\leq 1.3$), $M^*$ = -19.43, $\\alpha$ = -1.46; for red galaxies ($B_T-R_T > 1.3$), $M^*$ = -19.25, $\\alpha$ = -0.73, which were obtained by Marzke \\& da Costa (1997). The sample, which only considers galaxies out to a maximum distance of 8000 \\kms, contains 387 blue and 219 red galaxies. The results of the two-point correlation function are shown in Figure 13 (a) for redshift space while the fit parameters may be found in Table 8. Because of the small number of objects, the correlation function is very noisy, yet it is unquestionable that the red galaxies present a systematically higher amplitude at all separations compared to blue galaxies. In order to verify how sensitive the results may be to incompleteness, we re-calculated \\xis for the $m_B$=14.2 sample which is 92 \\% complete in colors. The fit parameters present a similar behavior, although the values differ from those measured for the 14.5 sample. The results we obtain for the samples discriminated in colors present a qualitative agreement with those of Tucker \\etal (1996), in the sense that red galaxies are more strongly correlated than blue galaxies. We have also calculated the real-space correlation function for the 14.5 sample and the power-law fit is presented in Fig. 13 (b), together with the redshift space correlation. The figure shows that the slopes of both power law fits are fairly similar ($\\gamma_r$=1.99 for blue, $\\gamma_r$=2.18 for red galaxies), though the uncertainties are rather large, in particular for the red galaxies. The observed \\xir suggests that red galaxies are probably more affected by peculiar motions than blue galaxies. Because of the relatively small size of the sample with colors, we have not been able to investigate the dependence on luminosity, which would be dominated by errors because of the small number of objects assigned to each luminosity bin. The relative bias estimated from $\\sigma_8$ in real space is $b_R$/$b_B$ = 1.40\\mm 0.33, and a similar result is obtained if the redshift space results are considered. As in the case of luminosity and morphology, one may calculate the relative bias between galaxies of different colors as a function of scale, which is presented in Fig. 14. Because the observed correlation function is rather noisy, for this plot we used the fits to \\xir . Taking the results at face value they would suggest that the relative bias between red and blue galaxies on small scales is comparable to that seen for early and late type galaxies. However, it levels off more rapidly ($\\sim $ 4 \\h1 Mpc), remaining constant at $b_R/b_B \\sim 1.2$ thereafter. This behavior could be the result of evolution due to environmental effects, where early type galaxies in higher density regions lost their gas more rapidly than bluer galaxies, and thus present a much lower star formation rate. However, because the errors are large, these results should only be considered as tentative. ", "conclusions": "" }, "9803/astro-ph9803083_arXiv.txt": { "abstract": "We review our present knowledge of high-redshift galaxies, emphasizing particularly their physical properties and the ways in which they relate to present-day galaxies. We also present a catalogue of photometric redshifts of galaxies in the Hubble Deep Field and discuss the possibilities that this kind of study offers to complete the standard spectroscopically based surveys. ", "introduction": "For a long time models for galaxy formation and evolution advanced unhampered by observations. Nowadays, however, the rapid increase in both observational capabilities and efficiency of the selection methods (see Steidel {\\em et al.} 1995 [S95]) has converted the task of looking for distant galaxies from one of the most difficult challenges to an almost routine job, and large databases of high-$z$ galaxies are already being compiled (Dickinson 1998, this Volume). Observations can now constrain the models, and this obliges us to understand the properties of these objects in order to get a complete image of the processes involved in the formation and evolution of galaxies. This study of the properties of high-$z$ galaxies is twofold. We need to understand the information provided by the confirmed high-$z$ galaxies. In this way we will learn about the spectral and morphological properties of the bright end of the galaxy population, i.e., the putative progenitors of present-day large ($L>L_*$) galaxies. Second, the use of photometric redshift techniques applied to deep multi-colour images (like the HDF, Williams {\\em et al.} 1996) opens a wealth of statistical methods to study those faint objects for which we cannot obtain spectroscopic information in the near future. These studies will yield further results on the general distribution and evolution of galaxies. The main problem for both methods resides in the $z \\approx 1-2$ range, where spectroscopic identification of galaxies at optical wavelengths is made difficult by the lack of spectral features. ", "conclusions": "The available data allow for different interpretations. While S95, S96 and G96 support the hypothesis that the observed high-$z$ galaxies are the progenitors of present-day luminous galaxies at the epoch of formation of the first stars in their spheroidal components, T97 suggests that these objects will evolve to form the Population II components of early-type spirals. Another interpretation (L97) maintains that these objects represent a range of physical processes and stages of galaxy formation and evolution rather than any particular class of object. While this third interpretation might be closer to reality, we are still missing an important piece of the puzzle. Detailed IR imaging and spectroscopy is needed in order to: a) shed light on the $z=1-2$ galaxies allowing us to constrain evolutionary models; b) obtain images of the $z>2$ galaxies at optical rest-frame wavelengths to be compared with their low-$z$ counterparts and; c) perform moderate resolution spectroscopy of the $z>2$ galaxies to accurately measure their metallicities and the importance of dust corrections. We expect that these observations, with the support of techniques like cosmological simulations and stellar population evolutionary models, will lead us closer to the long-searched-for understanding of the process by which the Universe came to be as we see it. Perhaps it is not the moment for us to ``look deeper in the Southern Sky'', but to look at it with different eyes." }, "9803/astro-ph9803340_arXiv.txt": { "abstract": "We examine the non-linear stability of the Wisdom-Holman (WH) symplectic mapping applied to the integration of perturbed, highly eccentric ($e\\gtrsim 0.9$) two-body orbits. We find that the method is unstable and introduces artificial chaos into the computed trajectories for this class of problems, {\\it unless} the step size is chosen small enough to always resolve periapse, in which case the method is generically stable. This `radial orbit instability' persists even for weakly perturbed systems. Using the Stark problem as a fiducial test case, we investigate the dynamical origin of this instability and show that the numerical chaos results from the overlap of step size resonances (cf. Wisdom \\& Holman 1992); interestingly, for the Stark problem many of these resonances appear to be absolutely stable. We similarly examine the robustness of several alternative integration methods: a regularized version of the WH mapping suggested by Mikkola (1997); the potential-splitting (PS) method of Lee et al. (1997); and two methods incorporating approximations based on Stark motion instead of Kepler motion (cf. \\cite{newet97}). The two fixed point problem and a related, more general problem are used to comparatively test the various methods for several types of motion. Among the tested algorithms, the regularized WH mapping is clearly the most efficient and stable method of integrating eccentric, nearly-Keplerian orbits in the absence of close encounters. For test particles subject to both high eccentricities and very close encounters, we find an enhanced version of the PS method---incorporating time regularization, force-center switching, and an improved kernel function---to be both economical and highly versatile. We conclude that Stark-based methods are of marginal utility in $N$-body type integrations. Additional implications for the symplectic integration of $N$-body systems are discussed. ", "introduction": "\\label{sec_intro} Symplectic integration schemes have become increasingly popular tools for the numerical study of dynamical systems, a result of their often high efficiency as well as their typical long-term stability (see, e.g., \\cite{marps96} and the many references within). The Wisdom-Holman (WH) symplectic mapping in particular (\\cite{wish91}; cf. \\cite{kinyn91}) has been widely used in the context of Solar System dynamics. However, the fact that this and other symplectic methods are, by construction, ``finely tuned'' can make them susceptible to performance-degrading ailments (much as high-order methods offer little benefit if the motion is not sufficiently smooth), and the stability of these methods for arbitrary systems and initial conditions is not completely understood. It would be prudent, therefore, to exercise caution when applying such schemes to systems entering previously unexplored dynamical states, and to ensure that adequate preliminary testing is undertaken regardless of the method's stability in previously considered problems. Recent galactic dynamics simulations by Rauch \\& Ingalls (1997), for example---which used the WH mapping---uncovered evidence of an instability in the method when applied to a particular class of problems: the integration of highly elliptical, nearly-Keplerian orbits in which the timestep is taken small enough to smoothly resolve the perturbation forces, but not so tiny as to explicitly resolve pericenter. Since particle motion in these simulations was extremely close to Keplerian near pericenter, and since the mapping itself is exact for Keplerian motion, there is no {\\it a priori} reason why the method should have performed as poorly and unstably as was found. Recently several variations of the WH mapping have been proposed which aim to extend the range of applicability of the original method. The regularized WH mappings investigated by Mikkola (1997), for instance, appear promising in the context of elliptical motion. The potential-splitting (PS) method of Lee, Duncan, \\& Levison (1997) allows symplectic integration of close encounters between massive bodies by adding a multiple-timestep algorithm similar to that of Skeel \\& Biesiadecki (1994). Unfortunately both of these schemes have limitations of their own; the former is unable to resolve close encounters, while the latter approach (like the original mapping) appears to be unstable when orbits are eccentric (cf. \\cite{dunll97}). In this paper, we use a series of test problems based on perturbed two-body motion to analyze the stability of the WH mapping and several of its variants. In particular, we examine the reliability of the methods for test particles whose motion is either highly eccentric or subject to close encounters with the perturbers (or both). The plan of the paper is as follows. In the following section, the performance of the WH mapping at high eccentricities is investigated using the Stark problem (e.g., \\cite{dan94}; \\cite{kir71})---for which the range of orbital eccentricities is easily controlled, and no close encounters occur---as the fiducial test case. The instability found in the integrated motion is then explained using complementary analytic and geometric arguments. In \\S~\\ref{sec_modwh}, modified forms of the original mapping are described; similarly, in \\S~\\ref{sec_sint} integrators based on Stark motion instead of Kepler motion are considered. Section~\\ref{sec_compsim} uses the two fixed point problem (e.g., \\cite{par65}) as well as a more general test problem (drawn from the area of galactic dynamics) to conduct a comparative performance analysis of the various algorithms. Both the Stark and two fixed point problems are fully integrable and analytically soluble in terms of elliptic functions and integrals, allowing a detailed assessment of the accuracy of the numerical results to be made. Concluding discussion is given in \\S~\\ref{sec_discuss}. ", "conclusions": "\\label{sec_discuss} We have shown that the WH mapping is generically unstable when applied to eccentric, nearly-Keplerian orbits whenever the step size is not small enough to resolve periapse. This `radial orbit instability' is fully explainable in terms of the overlap of step size resonances and has a simple geometric manifestation in the case of the Stark problem. Our investigation indicates that the islands of stability found in the latter problem do not exist in the more general cases we have examined; the instability therefore appears to be unavoidable in typical situations, unless one employs the brute-force approach of decreasing the timestep by the requisite amount. However, besides the fact that this is an extremely inefficient solution---it reduces the mapping to a very costly direct integration scheme---we have shown that an elegant solution to the problem is already available: the regularization approach of Mikkola (1997). In every case examined, not only was the regularized WH mapping immune to the radial orbit instability, in many cases it was also more efficient. We enthusiastically recommend its use whenever close encounters with perturbers are not of concern. We remind the reader that our investigation has not cast into doubt all previous studies that have used the WH integrator and its variants. In nearly every case care has been taken to use a small enough step size such that perihelion passage would be adequately sampled. We note only one area where particular caution should be exercised. One of the features of the long-term dynamics in mean motion resonances and secular resonance is that very high eccentricities can be developed. These eccentricities are often large enough that physical collision with the sun is a common outcome in studies of meteorite delivery from the main asteroid belt and the long-term dynamics of ecliptic comets (\\cite{glaet97}; \\cite{morm95}; \\cite{levd97}). In those cases it is unlikely that the step size used was small enough to resolve the perihelion passage. Although these researchers checked their results for step size dependence and reported no numerical artifacts, we suggest that further examination of those cases would be prudent. We have demonstrated that the potential-splitting method of Lee et al. (1997) can be regularized to produce an algorithm that is robust in the face of both close encounters and highly eccentric orbits. We have also shown that force-center switching during exceptionally close encounters can be cleanly incorporated into the method and can substantially enhance the stability of the algorithm without noticeably affecting its desirable symplectic qualities. We have not, however, found a practical way to regularize around the perturber while the switch is in effect; the stability of this approach during highly eccentric {\\it encounters} is correspondingly questionable. Our examination of Stark-based integrators indicates that they, too, are subject to the radial orbit instability unless regularized, although it tends to be less severe since the Stark approximation becomes systematically better near the origin. Unless the perturbing potential is {\\it very} well represented by a Stark potential, they also appear uncompetitive in terms of efficiency---by over an order of magnitude---due to the cost of Stark steps relative to that of Kepler steps. Among the Stark-based methods, the regularized, symplectic method (\\S~\\ref{sec_rss}) consistently outperformed the time-reversible method (\\S~\\ref{sec_trs}), in part because of the linearly growing energy error exhibited by the latter. Our conclusion is that Stark-based schemes are of marginal utility in the integration of N-body systems. It is clear that integrators based on a two fixed point (TFP) splitting instead of a Stark or Kepler splitting are also possible; they can be constructed in the same manner as the Stark-based methods were. Such methods could be useful whenever two bodies strongly dominate the mass in the system (e.g., asteroid motion in the Sun-Jupiter system). As for Stark motion, however, the relative expense of advancing the TFP Hamiltonian is a significant handicap, and the circumstances in which its use is justified remain unclear. On the other hand, since Stark motion is a subset of TFP motion it is not unlikely that methods based on the latter splitting will generically outperform those of the former type, since their analytic solutions are of similar complexity. It would be interesting to investigate this possibility in greater detail. Although we have confined attention to the perturbed two-body problem, the techniques employed in this paper are also applicable to general hierarchical $N$-body systems. In particular, we believe that regularization of the $N$-body version of the potential-splitting method (\\cite{dunll97}) is likely to cure the instability at high eccentricities noted by the authors. In principle force-center switching of the kind described in \\S~\\ref{sec_ps} can also be done, but we have not studied this possibility in detail. In any event, we have found the combination of regularization and potential-splitting to be a powerful one, and to produce a remarkably versatile symplectic method for the integration of nearly-Keplerian systems." }, "9803/astro-ph9803263_arXiv.txt": { "abstract": " ", "introduction": "The discovery of strong and remarkably coherent high-frequency X-ray brightness oscillations in at least sixteen neutron stars in low-mass binary systems has provided valuable new information about these stars, some of which are likely to become millisecond pulsars. Oscillations are observed both in the persistent X-ray emission and during thermonuclear X-ray bursts (see van der Klis 1997). The kilohertz quasi-periodic oscillations (QPOs) observed in the persistent emission have frequencies in the range 325--1200~Hz, amplitudes as high as $\\sim15$\\%, and quality factors $\\nu/\\delta\\nu$ as high as $\\sim200$. Two kilohertz QPOs are commonly observed simultaneously in a given source (see Fig.~1). Although the frequencies of the two QPOs vary by hundreds of Hertz, the frequency separation $\\Delta\\nu$ between them appears to be nearly constant in almost all cases (see van der Klis et al.\\ 1997 and M\\'endez et al.\\ 1997). \\begin{figure}[t] % \\begin{minipage}[b]{3.3in} \\centerline{ \\psfig{file=tokyo.fig1.eps,height=7.5truecm}} \\end{minipage} \\begin{minipage}[b]{2.3in} \\caption{Power density spectrum of Sco~X-1 brightness variations, showing the two simultaneous kilohertz QPOs that are characteristic. These are two of the weakest kilohertz QPOs observed, with rms amplitudes $\\sim1$\\%. The continuum power density is consistent with that expected from photon counting noise. From van der Klis et al.\\ (1997).} \\end{minipage} \\end{figure} The $\\sim$250--600~Hz brightness oscillations observed during type~I X-ray bursts are different in character from the QPOs observed in the persistent emission (see Strohmayer, Zhang, \\& Swank 1997). Only a single oscillation has been observed during X-ray bursts, and the oscillations in the tails of bursts appear to be highly coherent (see, e.g., Smith, Morgan, \\& Bradt 1997), with frequencies that are always the same for a given source (comparison of burst oscillations from \\fu{1728$-$34} over about a year shows that the timescale for any variation in the oscillation frequency is $\\gta 3000$~yr; Strohmayer 1997). The burst oscillations in \\fu{1728$-$34} and \\fu{1702$-$42} (see Strohmayer, Swank, \\& Zhang 1998) have frequencies that are consistent with the separation frequencies of their kilohertz QPO pairs. The burst oscillations in \\fu{1636$-$536} (Zhang et al.\\ 1997) and \\ks{1731$-$260} (Smith et al.\\ 1997) have frequencies that are consistent with twice the separation frequencies of their kilohertz QPO pairs (Zhang et al.\\ 1997; Wijnands \\& van der Klis 1997). The evidence is compelling that the burst oscillations are produced by rotation with the star of one or two nearly identical emitting spots on the surface (see Strohmayer et al.\\ 1997). The frequencies of the burst oscillations are therefore the stellar spin frequency or its first overtone. The frequency separation $\\Delta\\nu$ between the two kilohertz QPOs observed in the persistent emission of a given star is closely equal to the spin frequency of the star inferred from its burst oscillations (see Miller, Lamb, \\& Psaltis 1998, hereafter MLP). ", "conclusions": "" }, "9803/astro-ph9803055_arXiv.txt": { "abstract": " ", "introduction": "During the first thousand seconds in the evolution of the Universe, as it expanded and cooled from very high densities and temperatures, nuclear reactions transformed neutrons and protons into astrophysically interesting abundances of the light nuclides deuterium, helium-3, helium-4 and lithium-7. In the context of Standard, Big Bang Nucleosynthesis (SBBN; homogeneous, isotropic expansion, three flavors of non-degenerate neutrinos) these abundances depend on only one adjustable parameter, the nucleon density. Since as the Universe expands all densities decrease, it is useful to express the nucleon density in terms of a nearly constant parameter, the ratio of nucleons to photons, which has barely changed at all since the annihilation of electron-positron pairs in the early Universe. \\begin{equation} \\eta \\equiv n_{\\rm N}/n_{\\gamma} \\ \\ ; \\ \\ \\eta_{10} \\equiv 10^{10}\\eta \\end{equation} The contribution of nucleons (baryons) to the universal mass-density may be written as the dimensionless ratio of the baryon density to the critical density (which depends on the present value of the Hubble parameter: H$_{0} = 100\\,h\\,$kms$^{-1}$Mpc$^{-1}$; $\\Omega_{\\rm B} \\equiv \\rho_{\\rm B}/\\rho_{crit}$). \\begin{equation} \\Omega_{\\rm B}\\,h^{2} = \\eta_{10}/273 \\end{equation} SBBN is an overdetermined theory in that the observable abundances of four nuclides are predicted on the basis of one free parameter. In Figure 1 the predictions of the primordial abundances are shown for a wide range of $\\eta$. SBBN is falsifiable in that it is possible that {\\bf no} value of $\\eta$ will be consistent with the primordial abundances inferred from the observational data. Furthermore, consistency requires that {\\bf if} an acceptable value of $\\eta$ is found, the corresponding nucleon density at present, $\\Omega_{\\rm B}$, is in agreement with other astronomical observations. Indeed, since there must be enough baryons to account for the visible matter in the Universe, but not too many to violate constraints on the total mass density, the {\\it interesting} range of $\\eta$ in Figure 1 is restricted to $3\\times 10^{-11} - 1\\times 10^{-8}$. Even so, note the enormous range in the predicted abundances of deuterium and lithium. Over this same range in $\\eta$ the predicted primordial mass fraction of $^4$He, Y$_{\\rm P}$, hardly changes at all. As we shall soon see, consistency between D and $^4$He provides a key test of SBBN. \\begin{figure} \\centerline{\\psfig{file=fig1.ps,width=0.9\\textwidth}} \\vspace{-24pt} \\caption{SBBN-predicted abundances of the light nuclides versus $\\eta$. The $^4$He mass fraction (Y$_{\\rm P}$) is shown along with the ratio by number to hydrogen of D ($^2$H, $y_2$), $^3$He ($y_3$), and $^7$Li ($y_7$). This figure is from D. Thomas.} \\end{figure} \\subsection{Status Quo Ante} SBBN has provided one of the most spectacular confirmations of the standard, hot Big Bang model of cosmology. Along with the Hubble expansion and the cosmic background radiation, SBBN is one of the pillars of the standard model. It is the only one offerring a connection between particle physics and cosmology. For example, Walker \\etal (1991) reanalyzed the relevant observational data to make a critical confrontation between predictions and observations. Walker \\etal (1991) concluded that SBBN was consistent with the observational data for $\\eta_{10} = 3.4\\pm0.3$ ($\\Omega_{\\rm B}h^{2} \\approx 0.01$), making the nucleon density one of the very best determined of all cosmological parameters. Furthermore, they noted that to preserve this consistency required that the total number of ``equivalent\", light neutrinos (particles which were relativistic at BBN), N$_{\\nu}$, should not exceed 3.4. With the three known flavors of neutrinos (provided none has a mass comparable to MeV energies), this leaves very little room for any new (light) particles ``beyond the standard model\". At this point it may have been tempting to declare victory for SBBN and to move on to other problems in cosmology. However, it was still important to subject the standard model to ever more precise observational tests in order to reaffirm its consistency and to narrow even further the bounds on the nucleon density and on particle physics beyond the standard model. To our surprise, my colleagues and I found a dark cloud looming on the horizon of the standard model (Hata \\etal 1995). \\subsection{A Crisis For SBBN?} There had, in fact, always been a ``tension\" between the predictions of SBBN and the inferred primordial abundances of D and $^4$He (Kernan \\& Krauss 1994, Olive \\& Steigman 1995) in the sense that while deuterium favored ``high\" values of $\\eta$ (Steigman \\& Tosi 1992, 1995), helium-4 pointed towards lower values (Olive \\& Steigman 1995). Indeed, in a reanalysis focusing on the $^4$He abundance, Olive \\& Steigman (1995) found for the best estimate of the number of equivalent light neutrinos, N$_{\\nu} = 2.2$. Only a generous error estimate permitted consistency with SBBN. It was, therefore, not entirely unexpected when Hata \\etal (1995) identified a ``crisis\" for SBBN in their comparison of the best estimates of the primordial abundances derived from the observational data with those predicted by SBBN. The problem is illustrated in Figure 2 which concentrates on the key nuclides, D, $^4$He and $^7$Li. While the $^4$He abundance is just barely consistent with the low end of the $\\eta$ range identified by Walker \\etal (1991), the deuterium abundance is only consistent with the upper end of that range. Note that due to its ``valley\" shape and to the relatively larger uncertainties in its predicted and inferred abundances, lithium is consistent with either deuterium or helium. Since it thus fails to discriminate between the low $\\eta$ favored by helium and the higher $\\eta$ preferred by deuterium, lithium is ignored in the following discussion. \\begin{figure} \\centerline{\\psfig{file=fig2.ps,width=0.9\\textwidth}} \\caption{SBBN predictions (solid lines) for $^4$He (Y), D ($y_2$), and $^7$Li ($y_7$) with the theoretical uncertainties (1$\\sigma$) estimated by the Monte Carlo method (dashed lines). Also shown are the regions constrained by the observations at 68\\% and 95\\% C.L. (shaded regions and dotted lines, respectively). This figure is from Hata \\etal (1995).} \\end{figure} Three possible resolutions of the challenge to SBBN posed by the D -- $^4$He conflict suggest themselves. Perhaps the primordial abundance of helium inferred from observations of extragalactic \\hii regions (see, \\eg, Olive \\& Steigman 1995 and Olive, Skillman, \\& Steigman 1997) is too small (see, \\eg, Izotov, Thuan, \\& Lipovetsky 1994 and Izotov \\& Thuan 1997). If the primordial helium mass fraction were closer to 0.25 than to 0.23, the challenge to SBBN evaporates. Since several dozen \\hii regions are observed, the statistical uncertainty in Y is small, typically $\\pm 0.003$ or smaller (Olive, Skillman, \\& Steigman 1997, Izotov \\& Thuan 1997). But systematic errors, such as those due to uncertainties in the corrections for unseen neutral helium, for collisional ionization, for temperature fluctuations and, especially, for underlying stellar absorption, may well be much larger. Alternatively, it could be that our adopted primordial deuterium abundance is too small. If the true primordial ratio (by number) of deuterium to hydrogen were a few parts in $10^4$ rather than the few parts in $10^5$ inferred from observations in the solar system and the local interstellar medium (ISM), lower $\\eta$, consistent with Y$_{\\rm P}$, is allowed (see Fig. 2). This local estimate of the deuterium abundance requires an extrapolation from ``here and now\" (solar system, ISM) to ``there and then\" (primordial). Any errors in this extrapolation open the door to systematic errors. Finally, the possibility remains that our estimates of the primordial abundances are correct and the D -- $^4$He tension is a hint of ``new physics\". For example, if the tau neutrino were massive ($\\sim 5 - 20$ MeV) and unstable (lifetime $\\sim 0.1 - 10$ sec.), the ``effective\" number of equivalent light neutrinos would be less than the standard model case of N$_{\\nu}$ = 3 (Kawasaki \\etal 1994). For N$_{\\nu} = 2.1 \\pm 0.3$, consistency among the primordial abundances may be reestablished (Hata \\etal 1995, Kawasaki, Kohri, \\& Sato 1997). Other, non-standard, particle physics solutions are conceivable; degenerate neutrinos offer one such option (Kohri, Kawasaki, \\& Sato 1997). ", "conclusions": "The predictions of SBBN are observationally challenged. The primordial abundances of D and $^4$He inferred from observational data appear to be inconsistent with the predictions of SBBN. Several options present themselves with the potential to resolve this crisis. Perhaps the data are at fault. The conflicting deuterium abundances derived from observations of high-redshift, low-metallicity QSO absorbers point an incriminating finger. If these data are supplemented with solar system and ISM deuterium abundances, the lower D/H ratios are preferred. But, is the extrapolation from here and now (solar system, ISM) to there and then (primordial) under control, or might there be unidentified systematic errors lurking? The two sets of apparently inconsistent helium abundances suggest systematic errors at play in the extragalactic \\hii region abundance determinations. Although new data is always welcome, it is clear that a better understanding of existing data may prove even more important. In the absence of new data and/or a better understanding of the extant data it may be worthwhile to look elsewhere for clues. My colleagues and I (Steigman, Hata, \\& Felten 1997; SHF) have discarded the constraint on $\\eta$ from SBBN and have utilized four other observational constraints (Hubble parameter, age of the Universe, cluster gas (baryon) fraction, and effective ``shape\" parameter $\\Gamma$) to predict the three key cosmological parameters (Hubble parameter, total matter density, and the baryon density or $\\eta$). Considering both open and flat CDM models and flat $\\Lambda$CDM models, SHF tested goodness of fit and drew confidence regions by the $\\Delta\\chi^2$ method. In all of these models SHF find that large $\\eta_{10}$ ($\\gsim~6$) is favored strongly over small $\\eta_{10}$ ($\\lsim~2$), supporting reports of low deuterium abundances on some QSO lines of sight, and suggesting that observational determinations of primordial $^4$He may be contaminated by systematic errors." }, "9803/astro-ph9803325_arXiv.txt": { "abstract": "We present the results of a combined study of ASCA and ROSAT observations of the distant cluster Abell 2390. For this cluster a gravitational arc as well as weak lensing shear have been previously discovered. We determine the surface brightness profile and the gas density distribution of the cluster from the ROSAT PSPC and HRI data. A combined spatially resolved spectral analysis of the ASCA and ROSAT data show that the temperature distribution of the intracluster medium of A2390 is consistent with an isothermal temperature distribution in the range 9 to 12 keV except for the central region. Within a radius of $160 h_{50}^{-1}$ kpc the cooling time is found to be shorter than the Hubble time, implying the presence of a cooling flow. In this central region we find strong evidence for a multi-temperature structure. Detailed analysis of the combined ASCA and ROSAT data yields a self-consistent result for the spectral structure and the surface brightness profile of the cluster with a cooling flow of about $500 - 700$ M$_{\\odot}$ y$^{-1}$ and an age of about $10^{10}$ y. From the constraints on the temperature and density profile of the intracluster gas we determine the gravitational mass profile of the cluster and find a mass of about $2\\cdot 10^{15}$ M$_{\\odot}$ within a radius of $3 h_{50}^{-1}$ Mpc. A comparison of the projected mass profiles of the cluster shows an excellent agreement between the mass determined from X-ray data and the mass determined from the models for the gravitational arc and the weak lensing results. This agreement in this object, as compared to other cases where a larger lensing mass was implied, may probably be due to the fact that A2390 is more relaxed than most other cases for which gravitational lensing mass and X-ray mass have been compared so far. ", "introduction": "A 2390 is one of the most prominent clusters in the redshift range around z = 0.2. It was classified by Abell (1958) and Abell, Corwin \\& Olowin (1989) only as a richness class 1 cluster. As a target of the CNOC survey (Yee \\et\\ 1996a) deep photometric and spectroscopic data were obtained for this cluster, and these authors conclude that it should more likely be classified as a richness class 3 cluster (Yee \\et\\ 1996b). This new CNOC data also provide a mean redshift for this cluster of z=0.228 (compared to the previous literature value of z=0.232 by Le Borgne \\et\\ 1991). In X-rays it is among the ten brightest galaxy clusters known at a redshift larger than 0.18 (e.g. Ebeling \\et\\ 1996). It has been observed with the EINSTEIN observatory and showed a luminosity of $L_x \\sim 1.6 \\cdot 10^{45}$ \\egs (in the 0.7 to 3.5 keV energy band) (Ulmer \\et\\ 1986; if their result is converted to $H_0 = 50$ km s$^{-1}$ Mpc $^{-1}$ as used in this paper) and the cluster has a slightly elongated shape (McMillian \\et\\ 1989). A ``straight arc'' and several arclets were discovered in this clusters by Pello \\et\\ (1991). All these observations underline that A2390 is a very rich and massive cluster of galaxies. Pierre \\et\\ (1996) have analyzed a deep ROSAT HRI observation and found that the X-ray emission from A2390 is very concentrated and highly peaked, indicating a strong cooling flow of about 880 \\msu y$^{-1}$. The cooling flow is centered on the giant elliptical galaxy in the cluster center. This together with the observation of the strong lensing features may indicate that the very peaked central surface brightness is probably the effect of the cooling flow as well as of a steep central gravitational potential in the cluster. The straight arc has been modeled by Kassiola, Kovner, \\& Blandford (1992) and Narashima \\& Chitre (1993). Pierre \\et\\ (1996) have also modeled the lensing cluster with an elliptical potential model and a second clump in close consistency with the X-ray morphology. They find a projected mass within the arc radius of $M(r \\le 38'') \\sim 0.8 \\cdot 10^{14} h^{-1}$ \\msu. They compared the lensing mass with the X-ray data by taking the mass profile of the lensing model and the gas density profile from the X-ray surface brightness, and calculated the temperature profile needed to satisfy the hydrostatic equation. The bulk temperatures found by this approach are in the range 8 to 10 keV. This high temperature is consistent with the large X-ray luminosity of the cluster given the generally good correlation between X-ray ICM temperature and X-ray luminosity (e.g. Edge \\& Stewart 1991). A weak lensing shear in A2390 was also observed recently by Squires \\et (1996). They deduced an elliptical mass distribution, elongated in the direction of the straight arc. This is qualitatively consistent with the X-ray surface brightness distribution. As we will show in this paper the mass distribution inferred from the X-ray results are in excellent agreement with both the mass deduced from the weak lensing analysis and that from the strong lensing modeling. The detection of diffuse intracluster light was reported by V\\'ilchez-G\\'omez \\et\\ (1994). This may also be taken as a sign that the core of the cluster is relaxed and that the debris of tidal stripping of galactic halo material had enough time to settle in the gravitational potential of the cluster. All these previous studies and the fact that the cluster is very massive and X-ray luminous makes A2390 a perfect target for more detailed X-ray observations, in particular to compare a more precise mass determination from the X-ray data with the optical and lensing results. The indication of the fair agreement of the mass in the different previous studies and the existence of the strong cooling flow suggest that the cluster is essentially relaxed and therefore ideal for the test of the various methods of mass determination. In this paper we present a combined analysis of deep ASCA and ROSAT PSPC and HRI observations of this cluster. The ROSAT observations are discussed in Section 2 and the ASCA observations in Section 3. A combined analysis of the spectral data of both instruments which provides very interesting evidence for multi-temperature structure in the cooling flow region of A2390 is presented in Section 4. Section 5 contains the results of the cluster mass determination from the X-ray data, and compared to the lensing results in Section 6. The cooling flow structure is discussed in detail in Section 7. Section 8 provides a summary and conclusions. We use a value of $H_0 = 50$ km s$^{-1}$ Mpc$^{-1}$ for the Hubble constant throughout this paper. Thus 1 arcmin at the distance of A2390 corresponds to a scale of 277 $h_{50}^{-1}$ kpc. ", "conclusions": "The present study of A2390 with combined use of the ROSAT PSPC and ASCA GIS data yields important new results. The two most striking results of the current study are (1) good consistency between the cooling flow rates derived independently from the spectral and imaging analyses, and (2) excellent agreement between the total mass values determined from the X-ray data and the gravitational lensing. As outlined in the introduction, the cluster shows all the apparant features of a well relaxed cluster: elliptical symmetry, strong central concentration (and a cD galaxy which may be the result of this), a cooling flow, and a large intracluster light halo around the central galaxy. A long cooling time ($\\sim 10^{10}$ y) implied from the cooling flow rate may be taken as a reinforcement of the picture that the cluster was left quite undisturbed for a long time. In the morphological analysis of the HRI image of A2390 Pierre et al. (1996) found some indication of substructure in the cluster. They interpreted the substructural feature as a trace of substructure in the cluster potential, and such an excess potential is actually needed in the gravitational lensing model producing the observed gravitational arc. The subclump has an X-ray luminosity of only about 1/60 of that of the whole cluster and therefore its mass is less than about 1/15 of that of the cluster as concluded by Pierre et al. (1996). Such a small infalling mass component will not cause a significant disturbance on the equilibrium configuration of the cluster, and also may not influence the evolution of the cooling flow seriously. Therefore, the presence of this small substructure is not in contradiction to our finding that the cluster is generally well settled. The large measured iron abundance of $\\sim 0.3$ of the solar value is in line with other observed results for very rich nearby clusters and some distant clusters. We should note, however, that lower abundances have been found in some rich distant clusters like CL0016+16 (Furuzawa et al. 1997) and A851 (Mushotzky \\& Loewenstein, 1997, Schindler et al. 1998). Earlier studies have pointed out cases of striking differences between the lensing mass and X-ray determined mass (e.g. Miralda-Escud\\'e \\& Babul 1995). The reason for an excellent agreement between the two in A2390 is most probably found in that this cluster is fairly relaxed, as compared to other clusters studied by weak lensing technique and X-ray observations. For example, A2218 and A2163 show signs of recent merging (Squires et al. 1996b, 1997). The detailed study of A2218 shows a tendency that the lensing mass is higher than the X-ray mass, while in A2163 the two mass values are well consistent with each other. PKS0745 which also shows a strong cooling flow and a gravitational arc (Allen et al. 1996) may be in a similar situation to A2390, and consistency between the lensing mass and X-ray mass could be found in this system, too. Allen et al. (1996) have already stressed that the agreement of the mass determination in PKS0745 is most probably the result of the cluster being well relaxed (see also recent work in Allen 1997)." }, "9803/astro-ph9803113_arXiv.txt": { "abstract": "Chevalier \\& Ilovaisky (1998) use {\\it Hipparcos} data to show that the X-ray binary systems LSI+61$^\\circ$ 303 and A0535+262 are a factor of ten closer (i.e. $d\\sim$ few hundred pc) than previously thought ($d\\sim2$kpc). We present high quality CCD spectra of the systems, and conclude that the spectral types, reddening and absolute magnitudes of these objects are strongly inconsistent with the closer distances. We propose that the {\\it Hipparcos} distances to these two systems are incorrect due to their relatively faint optical magnitudes. ", "introduction": "In a recent paper Chevalier \\& Ilovaisky (1998 - CI98) presented {\\it Hipparcos} distances to 17 massive X-ray binary systems. In particular they presented results that appeared to indicate that two systems (LSI +61$^\\circ$ 303, A0535+262) were up to a factor 10 closer than previously thought. This has profound implications for any models one constructs for these systems, for instance suggesting they need only contain white dwarfs rather than neutron stars to explain their X-ray luminosity. In this paper we use CCD spectra to redetermine the spectral type of LSI +61$^\\circ$ 303 and A0535+262. In both cases we show that the derived spectral types and reddenings are strongly consistent with normal Be stars at the distances previously ascribed to the systems, and not with the new, closer distances. Finally we discuss how the discrepancy between the {\\it Hipparcos} and our distances may be explained in terms of the faint nature of these particular sources. ", "conclusions": "We have shown that the {\\it Hipparcos} derived distances ($\\sim$ few hundred pc) to two Be/X-ray binary systems, LSI +61$^\\circ$ 303 and A0535+262 are inconsistent with the spectral types, reddenings and apparent magnitudes of the objects. There is strong evidence that the `traditional' distances to these objects (each $\\sim 2$kpc) are in fact correct. We note here that these two objects have the worst goodness-of-fit values in the {\\it Hipparcos} catalogue (ESA 1997) of the CI98 sample (although they do lie below the maximum ``acceptable'' value of 3). In addition they are the faintest in the sample. This appears to indicate that the application of the simple `goodness-of-fit' criterion that anything less than 3 is a good parallax to faint objects is not reliable, and that the interpretation of {\\it Hipparcos} parallax data should always be carried out with this in mind." }, "9803/astro-ph9803169_arXiv.txt": { "abstract": "We report results of $^{12}$CO ($J=1$-0) mapping observations of the Wolf-Rayet starburst galaxy Mrk 1259 which has optical evidence for the superwind seen from a nearly pole-on view. The CO emission is detected in the central 4 kpc region. The nuclear CO spectrum shows a blue-shifted ($\\Delta V \\simeq -27$ km s$^{-1}$) broad (FWHM $\\simeq$ 114 km s$^{-1}$) component as well as the narrow one (FWHM $\\simeq 68$ km s$^{-1}$). The off-nuclear CO spectra also show the single-peaked broad component (FWHM $\\simeq$ 100 km s$^{-1}$). The single-peaked CO profiles of both the nuclear and off-nuclear regions may be explained if we introduce a CO gas disk with a velocity dispersion of $\\sim 100$ km s$^{-1}$. If this gas disk would be extended up to a few kpc in radius, we may explain the wide line widths of the off-nuclear CO emission. Alternatively, we may attribute the off-nuclear CO emission to the gas associated with the superwind. However, if all the CO gas moves along the biconical surface of the superwind, the CO spectra would show double-peaked profiles. Hence, the single-peaked CO profiles of the off-nuclear regions may be explained by an idea that the morphology and/or velocity field of the molecular-gas superwind are more complex as suggested by hydrodynamical simulations. ", "introduction": "In starburst galaxies, a large number of massive stars (e.g., $\\sim 10^{4-5}$) are formed within a short duration (Weedman et al. 1981; Balzano 1983; Taniguchi et al. 1988). Therefore, a burst of supernova explosions occurs inevitably $\\sim 10^7$ years after the onset of the starburst. Since these numerous supernovae release a huge amount of kinetic energy into the circumnuclear gas, the circumnuclear gas is thermalized and then blow out into the direction perpendicular to the galactic disk as a ``superwind'' (Tomisaka \\& Ikeuchi 1988; Heckman, Armus, \\& Miley 1990; Suchkov et al. 1994). A bubble of the ionized gas sweeps up the circumnuclear molecular gas, leading to the formation of molecular-gas superwind as well as the ionized-gas one (Tomisaka \\& Ikeuchi 1988; Suchkov et al. 1994). Thus, in order to understand the whole physical processes of superwinds, it is important to investigate the nature of molecular-gas superwinds (e.g., Nakai et al. 1987; Aalto et al. 1994; Irwin \\& Sofue 1996). In this {\\it Letter}, we present new evidence for the molecular-gas superwind from the Wolf-Rayet starburst galaxy Mrk 1259, which shows the optical evidence for the superwind viewed from a nearly pole-on view (Ohyama, Taniguchi, \\& Terlevich 1997; hereafter Paper I). Mrk 1259 is a peculiar S0 galaxy (de Vaucouleurs et al. 1991; hereafter RC3) at a distance of 26.64 Mpc \\footnote{Paper I adopted a distance toward Mrk 1259, $D = 33.5$ Mpc. However, $V_{\\rm 3K}$ was misused instead of $V_{\\rm GSR}$ in this estimate. In this {\\it Letter}, using $V_{\\rm GSR} = 1998$ km s$^{-1}$ (RC3), with a Hubble constant $H_0$ = 75 km s$^{-1}$ Mpc $^{-1}$, we adopt a distance $D$ = 26.64 Mpc. Therefore, the HeII$\\lambda$4686 luminosity, the number of late WR (WRL) stars, the size of the superwind, and the average velocity of the superwind in Paper I should be read as $L$(HeII) = 7.0$\\times 10^{39}$ erg s$^{-1}$, $N$(WRL) $\\simeq$ 4100, $r$(superwind) $\\simeq 3.3$ kpc, and the average wind velocity $\\simeq$ 565 km s$^{-1}$, respectively.}. The logarithmic major-to-minor diameter ratio, log $R_{\\rm 25}=0.10\\pm 0.08$ (RC3), gives a nominal inclination angle, $i=37\\fdg 4^{+11.2}_{-20.1}$, and the galaxy appears to be elongated along the EW direction. If this elongation were attributed to the inclination, we would observe the rotational motion along the EW direction. However, our long slit optical spectrum along the EW direction which was analyzed in Paper I shows no hint on the rotational motion; $\\Delta V\\lesssim 50$ km s$^{-1}$, suggesting strongly that the galaxy is seen from an almost face-on view. Therefore the oval shape of Mrk 1259 may not be due to the inclination\\footnote{It seems no surprise even if an isolated galaxy shows some morphological peculiarity because any galaxy would experience some minor merger events in its life. It is also noted that minor mergers can cause nuclear starbursts (e.g., Hernquist \\& Mihos 1995; Taniguchi \\& Wada 1996).}. ", "conclusions": "In Table 1, we give a summary of our observational results. The integrated CO intensity was estimated by $I({\\rm CO}) = \\int T_{\\rm A}^* \\eta_{\\rm mb}^{-1} dv$ K km s$^{-1}$ where $\\eta_{\\rm mb} = 0.51$. Using a galactic conversion factor, $N_{\\rm H_{2}}/I_{\\rm CO} = 3.6\\times 10^{20}$ cm$^{-2}$ (K km s$^{-1}$)$^{-1}$ (Scoville et al. 1987), we estimate the molecular gas mass, $M_{\\rm H_2} = 5.8 \\times 10^6 I({\\rm CO}) A$, in each position where $A$ is the projected area of a 15$^{\\prime\\prime}$ HPBW in units of kpc$^2$. For the off-nuclear regions, we also give total values of $I$(CO) and $M_{\\rm H_2}$. The total molecular gas mass detected in our observations amounts to $1.2 \\times 10^9 M_\\odot$. Since we do not observe the entire disk of this galaxy, this mass is regarded as a lower limit. \\subsection{The Nuclear CO Emission} The nuclear CO emission shows a single-peaked profile with the evident blueward asymmetry. Applying a two-component Gaussian profile fitting (see the midst panel of Figure 1), we obtain the blueshifted broad component with FWHM $\\simeq$ 114 km s$^{-1}$ and the narrow one with FWHM $\\simeq$ 68 km s$^{-1}$. The peak velocity of the broad component is blueshifted by 27 km s$^{-1}$ with respect to that of the narrow one (Table 1). Both the intensities are nearly the same. We also mention that the red wing cannot be seen in the nuclear CO profile. Even though there is the broad CO emission component, its width is significantly narrower than those observed for typical starburst galaxies; e.g., FWHM(CO) $\\simeq$ 200 - 250 km s$^{-1}$ for M82 (Young \\& Scoville 1984; Nakai et al. 1987), $\\sim 350$ km s$^{-1}$ for NGC 1808 (Aalto et al. 1994), and $\\sim 325$ km s$^{-1}$ for NGC 4945 (Dahlem et al. 1993). This difference can be attributed to the effect of viewing angles between Mrk 1259 and the other starburst galaxies. It is remembered that the CO line width is generally affected by the galactic rotation. Given a typical rotation velocity of a disk galaxy, $V_{\\rm rot} \\sim 200$ km s$^{-1}$, the observed full widths would amount to 2$V_{\\rm rot} \\sim$ 400 km s$^{-1}$ if seen from the edge-on view. In fact, since we observe M82, NGC 1808, and NGC 4945 from highly inclined viewing angles, their line widths are considered to be broadened by the effect of galactic rotation. On the other hand, since Mrk 1259 appears to be a nearly face-on galaxy, the observed width is not affected by the galactic rotation. Irwin \\& Sofue (1996) suggested that one of the nearby superwind galaxies, NGC 3628, has a nuclear molecular gas disk with a velocity dispersion of $\\sim$ 100 km s$^{-1}$. If Mrk 1259 has also such a nuclear gas disk, we can explain the velocity width of the nuclear CO emission. Therefore, it is suggested that the observed FWHM of Mrk 1259 is due mainly to the broadening by some dynamical effect of the starburst activity. The blueward asymmetry of the nuclear CO line profile suggests that the CO gas is affected significantly by the superwind. \\subsection{The Off-Nuclear CO Emission} The detection of the off-nuclear CO emission from Mrk 1259 is very intriguing from the following two points. The first point is that the host galaxy of Mrk 1259 appears to be an S0 galaxy (RC3). It is often observed that early type galaxies such as S0 and elliptical galaxies tend to have less molecular gas (e.g., Young \\& Scoville 1991) although CO emission has been detected from a number of S0 galaxies (Thronson et al. 1989; Wiklind \\& Henkel 1989; Sage 1989; Sage \\& Wrobel 1989). It is also known that the molecular gas in (non-active) S0 galaxies tends to be concentrated in the region whose diameter is typically less than one tenth of the optical diameter (Taniguchi et al. 1994). If this is also the case for Mrk 1259, the molecular gas would be concentrated within the central $12\\arcsec =0.1 D_{\\rm 0}$ region where $D_{\\rm 0}$ is the isophotal optical diameter (RC3). Therefore, the presence of the bright off-nuclear CO emission is one of very important characteristics of Mrk 1259. The second point is that the line widths of the off-nuclear CO emission are comparable to that of the nuclear CO emission, FWHM $\\sim 100$ km s$^{-1}$. If there were an inclined off-nuclear CO disk, we may explain the wide line width because of the velocity gradient in the disk. If this is the case, we would observe that the peak velocity at 15$^{\\prime\\prime}$E is significantly different from that at 15$^{\\prime\\prime}$W. However, since our observations show that the velocity field of the off-nuclear regions is almost symmetric, this possibility is rejected. The second possibility is that there are spatially extended starburst regions and a significant amount of molecular gas is associated with them. However, radio continuum (1.5 GHz and 5 GHz) images show that the starburst region of Mrk 1259 is concentrated in the central several arcsec region (R. A. Sramek 1997, private communication). Therefore, there is no observational evidence for active star forming regions in the off-nuclear regions. As described before, we are observing the disk of Mrk 1259 from nearly a face-on view and thus the CO line width would be as narrow as $\\sim$ 10 km s$^{-1}$ if Mrk 1259 were a normal disk galaxy (Lewis 1984, 1987; Kamphuis \\& Sancisi 1993). If there were an extended molecular gas disk with a velocity dispersion of $\\sim$ 100 km s$^{-1}$ up to a radius of a few kpc, we could explain the wide line width. However, the size of the nuclear gas disk in NGC 3628 is much smaller ($\\simeq 230$ pc, or $\\sim 0.01 D_{\\rm 0}$) than the off-nuclear distance of Mrk 1259 ($\\sim 2$ kpc, or $\\sim 0.13 D_{\\rm 0}$). Although we cannot rule out the possibility that Mrk 1259 has such a very extended molecular gas disk with a large velocity dispersion, we need further detailed molecular-line observations to confirm this possibility. The third possibility is that the off-nuclear CO gas is associated with the superwind (i.e., blown out from the nuclear region). Since the ionized-gas superwind is extended to $r \\sim 3.3$ kpc (Paper I), this possibility seems to be quite high. In fact, such extended CO emission is detected in M82 at the scale of 600 pc (Nakai et al. 1987) and even at the larger scale ($\\sim 2$ kpc; Sofue et al. 1992). We discuss this possibility in detail in the next section. \\subsection{Biconical Superwind Model for Mrk 1259} Since the superwind of Mrk 1259 is observed from nearly the pole-on view, it is interesting to investigate both the velocity field and the geometry of the superwind. In order to perform this, we investigate the off-nuclear CO line profile using a simple biconical outflow model in which the superwind flows toward the polar directions symmetrically with its apex at the nucleus. Such a superwind geometry is expected theoretically by hydrodynamical numerical simulations (Suchkov et al. 1994) and indeed observed in M82 (e.g., Nakai et al. 1987). In our model, we assume that the molecular gas can only move along the cone surface. We assume that the axis of the cone lies along our line of sight. The full opening angle of the cone ($\\theta$) is not well constrained by the observations because of its nearly face-on viewing angle. Therefore we take this as a free parameter although Paper I has suggested as $\\theta \\lesssim 90\\arcdeg$. A mean tangential velocity of the ionized gas on the sky can be estimated as $V_{\\rm t, ion}\\simeq R_{\\rm SW}/T_{\\rm SW}\\simeq (2.3$ kpc$) /(5.5\\times 10^6$ years)$\\simeq 410$ km s$^{-1}$ where $R_{\\rm SW}$ is the projected radius of the superwind and $T_{\\rm SW}$ is the age of the superwind (Paper I). We note that the outflow velocity of the molecular gas is {\\it slower} than that of the ionized gas because the molecular gas along the cone surface is {\\it dragged} by the ionized gas, rather than directly {\\it pushed out} (Suchkov et al. 1994). For example, the model A1 of Suchkov et al. (1994) shows that the velocity of the outflowing dense gas is slower by a factor of $\\sim 5$ than that of the ionized gas at the age of 8.3 Myr. In fact, comparing the outflow velocity of the ionized gas (Heckathorn 1972) with that of the molecular gas (Nakai et al. 1987) of M82, we find that the outflow velocity of the molecular gas is slower by a factor of $\\sim 3$ than that of the ionized-gas. Thus, the mean tangential velocity of the molecular gas on the sky can be $V_{\\rm t, mol}=V_{\\rm t, ion}/\\epsilon$ where $\\epsilon$ is the decelerating factor ($\\epsilon \\simeq 3 - 5$). We examine if the model can explain the observed CO line profiles in the off-nuclear regions. No effect of radiative transfer is included in the model calculation. We assume that the size of the cone is large enough to cover the whole off-nuclear regions. For simplicity, we also assume that the velocity field has a power-law form; i.e., $V(r) \\propto r^a$, with a boundary condition of $V_{\\rm t, ion}$ ($r = 2.3$ kpc) = 410 km s$^{-1}$. The emissivity (strength of the CO emission per a unit area) is also assumed to have a power-law form; i.e., $I(r) \\propto r^b$. Although the parameters $a$ and $b$ are not well constrained by the observations, we adopt $a = 1$ and $b = -1$ as representative values following the trend seen in M82 (Nakai et al. 1987). We calculate the model for the cases of $\\theta = 60\\arcdeg, 90\\arcdeg, 120\\arcdeg$, and $150\\arcdeg$ and $\\epsilon$ = 1, 2, 3, 4, 5, and 6. To explain the observed FWZI (Full Width at Zero Intensity) of the off-nuclear CO emission ($\\sim 200$ km s$^{-1}$; see Figure 1), we find that only models with ($\\theta = 90\\arcdeg$ and $\\epsilon \\simeq 5 - 6$) and ($\\theta = 120\\arcdeg$ and $\\epsilon \\simeq 3 - 4$) are acceptable. Models with $\\theta = 60\\arcdeg$ and $150\\arcdeg$ cannot reproduce the line width for any $\\epsilon$. Therefore, we show our results only for the cases $\\theta=90\\arcdeg$ and $120\\arcdeg$ and $\\epsilon$ = 3, 4, 5, and 6 in Figure 2. Our simple model demonstrates that the CO line has always a double-peaked profile for any combinations of the parameters. The red peak corresponds to the recessing cone while the blue one corresponds to the advancing cone. On the other hand, our observations show that the off-nuclear CO lines have smooth profiles around at the systemic velocity. We discuss why our simple model cannot reproduce the observed off-nuclear CO profiles. One possible idea is a ``swirl''-like velocity field which is often found in the hydrodynamical numerical simulations (Tomisaka \\& Ikeuchi 1988; Suchkov et al. 1994). If the outflow actually shows such a complex geometry and/or velocity field, it is expected that some parts of the emission would contribute to the core emission and can explain the broad and smooth off-nuclear emission. In order to understand the molecular-gas superwind of Mrk 1259, detailed molecular-line observations with higher spatial resolution would be helpful. \\vspace{0.5cm} We would like to thank the staff of Nobeyama Radio Observatory for their kind support for our observations. We thank Naomasa Nakai for useful discussion and encouragement and Takashi Murayama and Shingo Nishiura for kind assistance of the observations. We also thank R. A. Sramek for kindly providing us his VLA data. YO was supported by the Grant-in-Aid for JSPS Fellows by the Ministry of Education, Culture, Sports, and Science. This work was financially supported in part by Grant-in-Aids for the Scientific Research (No. 0704405) of the Japanese Ministry of Education, Culture, Sports, and Science. \\newpage \\begin{table} \\caption{Molecular gas properties of Mrk 1259} \\begin{tabular}{llccc} \\tableline \\tableline & & & Nucleus & Off-nucleus \\\\ \\tableline rms noise & $\\delta T_{\\rm A}^*$ (K) & & 0.018 & $\\sim 0.015$ \\\\ Flux & $I$(CO)$^{a, b}$ (K km s$^{-1}$) & total & $28.0\\pm 1.1$ & $43.8\\pm 1.9$$^d$ \\\\ & & 15\\arcsec N & & $4.2\\pm 1.0$ \\\\ & & 15\\arcsec E & & $18.6\\pm 1.0$ \\\\ & & 15\\arcsec S & & $9.5\\pm 1.0$ \\\\ & & 15\\arcsec W & & $11.5\\pm 1.0$ \\\\ Mass & $M_{\\rm H_2}^{b, c}$ ($M_\\odot$) & & $(4.8\\pm 0.2) \\times 10^8$ & $(7.5\\pm 0.3) \\times 10^8$$^d$ \\\\ Line profile & & & & \\\\ & FWHM$_{\\rm narrow}$ (km s$^{-1}$) & & 68 & \\nodata \\\\ & $V_{\\rm narrow}$ (km s$^{-1}$) & & 2178 & \\nodata \\\\ & FWHM$_{\\rm broad}$ (km s$^{-1}$) & & 114 & $\\sim 100$$^e$ \\\\ & $V_{\\rm broad}$ (km s$^{-1}$) & & 2151 & $\\sim 2170$$^e$ \\\\ & $I$(broad)/$I$(narrow) & & 1.0 & \\nodata \\\\ \\tableline \\end{tabular} \\tablenotetext{a}{$I({\\rm CO}) = \\int T_{\\rm A}^* \\eta_{\\rm mb}^{-1} dv$ K km s$^{-1}$ where $\\eta_{\\rm mb}$ = 0.51.} \\tablenotetext{b}{Formal one-sigma error indicated.} \\tablenotetext{c}{A Galactic conversion factor, $N_{\\rm H_{2}}/I_{\\rm CO} = 3.6\\times 10^{20}$ cm$^{-2}$ (K km s$^{-1}$)$^{-1}$ (Scoville et al. 1987), is assumed.} \\tablenotetext{d}{Sum of the fluxes of all the four off-nuclear regions.} \\tablenotetext{e}{No profile fitting was made because of both the lower signal-to-noise ratio and the non-Gaussian profiles. The values shown here are typical ones for the off-nuclear regions except 15$^{\\prime\\prime}$N.} \\end{table} \\newpage" }, "9803/gr-qc9803068_arXiv.txt": { "abstract": "We propose two new classes of instantons which describe the tunneling and/or quantum creation of closed and open universes. The instantons leading to an open universe can be considered as generalizations of the Coleman-De-Luccia solution. They are non-singular, unlike the instantons recently studied by Hawking and Turok, whose prescription has the problem that the singularity is located on the hypersurface connecting to the Lorentzian region, which makes it difficult to remove. We argue that such singularities are harmless if they are located purely in the Euclidean region. We thus obtain new singular instantons leading to a closed universe; unlike the usual regular instantons used for this purpose, they do not require complex initial conditions. The singularity gives a boundary contribution to the action which is small for the instantons leading to sufficient inflation, but changes the sign of the action for small $\\phi$ corresponding to short periods of inflation. ", "introduction": "\\label{sec-nonsing} Suppose we have an effective potential $V(\\phi)$ with a local minimum at $\\phi_1$, and a global minimum at $\\phi=0$, where $V=0$ (see Fig. \\ref{potential}). In an $O(4)$-invariant Euclidean spacetime with the metric \\begin{equation}\\label{metric} ds^2 =d\\tau^2 +a^2(\\tau)(d \\psi^2+ \\sin^2 \\psi \\, d \\Omega_2^2) \\ , \\end{equation} the scalar field $\\phi$ and the three-sphere radius $a$ obey the equations of motion \\begin{equation}\\label{equations} \\phi''+3{a'\\over a}\\phi'=V_{,\\phi},~~~~~ a''= -{8\\pi G\\over 3} a ( \\phi'^2 +V) \\ , \\end{equation} where primes denote derivatives with respect to $\\tau$. \\begin{figure}[Fig0] \\hskip 1.5cm \\leavevmode\\epsfysize=4cm \\epsfbox{Potential.eps} \\ \\caption[Fig1]{\\label{potential} Effective potential $V(\\phi) = {m^2\\over 2}(\\phi^2(\\phi- v)^2 +B \\phi^4)$ for $m^2 = 2$, $B=0.12$ and $v = 0.5$. It has a shallow minimum at $\\phi_0 = 0.357$ and a local maximum at $\\phi_1=0.312$. All quantities in this figure are in units of $M_{\\rm p}/\\sqrt{8\\pi}$.} \\end{figure} These equations have several non-singular solutions, the simplest of which are the $O(5)$ invariant four-spheres one obtains when the field $\\phi$ sits at one of the extrema of its potential. In this case the first of the two equations above is trivially satisfied, and $a(\\tau) = H^{-1} \\sin H\\tau$. Here $H^2 = {8\\pi V\\over 3 M_{\\rm p}^2}$. Using the solution for which $\\phi=\\phi_1$, Hawking and Moss \\cite{HM} found the rate at which the field $\\phi$ in a single Hubble volume tunnels to the top of the potential, from which it can roll down towards the true vacuum. For a recent discussion of this instanton and its interpretation see \\cite{ALOpen}. The main other use of these trivial instantons is to find the action of the false vacuum background solution, which must be subtracted from the bounce action to obtain a tunneling rate. We shall consider potentials for which $V_{,\\phi\\phi} \\gg H^2$ in the region where the tunneling occurs. In this case, tunneling out of the false vacuum does not occur primarily on the scale of an entire Hubble volume via the Hawking-Moss instanton. Instead the transition will proceed via more complicated Euclidean solutions with varying field $\\phi$. These include the Coleman-De-Luccia instanton, and related instantons which we found. \\subsection{Bubble instantons} A Euclidean solution which describes the creation of an open universe was first found by Coleman and De Luccia in 1980 \\cite{CL}. It is given by a slightly distorted de~Sitter four-sphere of radius $H^{-1}(\\phi_0)$. Typically, the field $\\phi$ is very close to the false vacuum, $\\phi_0$, throughout the four-sphere except in a small region (whose center we may choose to lie at $\\tau=0$), in which it lies on the `true vacuum' side of the maximum of $V$. The behavior of the field and scale factor for the potential in Fig.~\\ref{potential} is shown in Fig.~\\ref{Colem}. The scale factor vanishes at the points $\\tau=0$ and $\\tau=\\tau_{\\rm f} \\approx \\pi/H$, which we will call the North and South pole of the four-sphere. In order to get a singularity-free solution, one must have $\\phi' = 0$ and $a'=\\pm 1$ on the poles. This solution can be cut in half along the line $\\psi=\\pi/2$, which removes half of each three-sphere. Then one can continue analytically to a Lorentzian spacetime~\\cite{GutWei83,HT} with the time variable $\\sigma$, given by $\\psi=\\pi/2+i\\sigma$. This gives region II of the Lorentzian universe (see Fig.~\\ref{fig-regions}): \\begin{equation} ds^2 = -a^2(\\tau)\\ d\\sigma^2 + d\\tau^2 + a^2(\\tau) \\cosh^2 \\sigma\\ d\\Omega_2^2; \\end{equation} the field $\\phi$ will still depend on $\\tau$ in the same way as before, and will be independent of $\\sigma$. This describes a shell of width $H^{-1}$, which is mostly near the false vacuum and expands exponentially. The shell separates two bubbles, regions I and III, in which the universe looks open. \\begin{figure}[Fig1] \\hskip 1.5cm \\leavevmode\\epsfysize=9.5cm \\epsfbox{Coleman.eps} \\ \\caption[Fig1]{\\label{Colem} The upper panel shows the behavior of the scalar field $\\phi$ for examples of the Coleman-De-Luccia ``bubble'' instanton (solid line) and the new ``double-bubble'' instanton which we have found (dashed line). For both instantons, the field is in the domain of the true vacuum at small $\\tau$, forming a bubble. For the bubble instanton, the field is closest to the false vacuum at the pole opposite the bubble. For the double-bubble instanton, this happens on the equator, at the moment of the maximal expansion. The behavior of the three-sphere radius $a(\\tau)$ shown in the lower panel is very similar for both instantons, though it is not identical.} \\end{figure} Region I is obtained by taking $\\sigma = i\\pi/2 + \\chi$ and $\\tau = it$, giving the metric \\begin{equation} ds^2 = -dt^2 + \\alpha^2(t) \\left( d\\chi^2 + \\sinh^2 \\chi d\\Omega_2^2 \\right), \\end{equation} where $ \\alpha(t) = -i\\, a[\\tau(t)]$. Its spacelike sections (defined by the hypersurfaces of constant inflaton field) are open. Thus, region I looks from the inside like an infinite open universe, which inflates while the field $\\phi$ slowly rolls down to the true vacuum. The evolution will then undergo a transition to a radiation or matter-dominated open Friedman-Robertson-Walker universe. In region III, which is obtained by choosing $\\sigma = i\\pi/2 + \\chi$ and $\\tau = \\tau_{\\rm f} + it$, the field $\\phi$ rolls to the local minimum at $\\phi_0$, and one gets indefinite open inflation in the false vacuum. \\begin{figure}[Fig1] \\hskip 1.5cm \\leavevmode\\epsfysize=5cm \\epsfbox{continuation.eps} \\ \\caption[Fig1]{\\label{fig-regions} The Lorentzian de~Sitter-like spacetime obtained from the analytic continuation of Coleman-De-Luccia instantons contains three regions. In Regions I and III the hypersurfaces of constant field $\\phi$ form open spacelike sections. Region II is a shell separating the two bubbles.} \\end{figure} The analytic continuations we have given support the interpretation of such solutions as the spontaneous nucleation of a bubble of true vacuum on the background of de~Sitter space expanding in the false vacuum. For this reason we will call them `bubble instantons'. The nucleation rate is given by \\begin{equation} \\Gamma = e^{-\\Delta S}, \\end{equation} where $\\Delta S$ is the difference between the action of the full Euclidean bubble solution, and the action of a Euclidean solution describing the background spacetime. Except for near-Planckian potentials, both actions will be large and negative (about $-2.6 \\times 10^4$ in our example). The background solution is given by an exact Euclidean four-sphere on which the field $\\phi$ is constant and equal to $\\phi_0$, the false vacuum. Its action will be $-{3 M_{\\rm p}^4 \\over 8 V(\\phi_0)}$. Subtracting this from the action of the bubble solution, one obtains a positive $\\Delta S$ ($\\approx 4.9$ in our example). This means that bubble formation by tunneling is suppressed, as it should be. One usually requires instanton solutions to interpolate between the initial and final spacelike sections (in this case, a section of pure de~Sitter space in the false vacuum and a similar section containing a bubble of true vacuum). The above description, which seems to use two disjoint instantons, is actually consistent with this formal requirement, since the instantons may be connected by virtual domain walls after small (Planck size) four-balls are removed. This will cause the background instanton to contribute to the total action with a negative sign. If one connects the background instanton to the region of the bubble instanton where $\\phi$ is closest to its false vacuum, the discontinuity in $\\phi$ will be small, so the volume contributions of the removed regions cancel almost exactly. Requiring continuous instantons, therefore, does not change the pair creation rate significantly~\\cite{BC}. Cosmological instantons have frequently been interpreted to describe the creation of a universe from nothing, i.e.\\ without a pre-existing background. This case is considerably less well-defined than the quantum nucleation of structures on a given background solution. In particular, the sign with which the large, negative action enters the exponent in the path integral is subject to debate~\\cite{HT,ALOpen,HTnew}. Leaving such questions aside for now, we will take the position that isolated cosmological instantons are indeed related to universe creation, independently of the formalism used to assign probabilities to such processes. \\subsection{Double-bubble instantons} We have found a new instanton in which there are two bubbles, one on each pole. In this solution, $\\phi$ is in the domain of the true vacuum in small regions near the poles, and near the false vacuum elsewhere; this can be seen from the dashed line in Fig.~\\ref{Colem}. The geometry is still approximately a four-sphere. As before, $\\phi'$ vanishes on the poles; but now it also vanishes on the equator, at $\\tau=\\tau_{\\rm max}$. The Northern and Southern hemispheres are exactly symmetric. Not surprisingly, the action of the double-bubble solution, after the background subtraction described above, is approximately twice that of the bubble (Coleman-De-Luccia) instanton. For the instanton shown in Fig.~\\ref{Colem} one has $\\Delta S_2 \\approx 9.8$. The analytic continuations will be the same as before, with a different result. Region II will be mostly in the domain of the false vacuum. Region I and III will be identical, each containing an open inflating universe in which the field rolls down to the true vacuum. Globally, therefore, we obtain two bubbles of true vacuum separated by a shell which inflates in the false vacuum. This solution can be interpreted as the spontaneous pair-creation of bubbles of open inflation on the background of false vacuum inflation. Alternatively, one may view it as the creation from nothing of two open inflating universes separated by a metastable shell. \\subsection{Anti-double-bubble instantons} In addition we have found another family of instantons, two examples of which are shown in Fig.~\\ref{loop}. In these instantons, the field is in the domain of the false vacuum in two regions surrounding the poles. They are separated by a thin shell at the equator, where the field is in the true vacuum domain. These instantons have a much greater action difference to the background instanton, since the true-vacuum region is significantly larger than in the previous two cases. In particular, $\\Delta S = 93.6$ for the instanton shown by the solid line in Fig.~\\ref{loop}, and $\\Delta S = 124.7$ for the instanton shown by the dashed line. \\begin{figure}[Fig0111] \\hskip 1.5cm \\leavevmode\\epsfysize=4.5cm \\epsfbox{NewInst.eps} \\ \\caption[Fig1]{\\label{loop} Two examples of ``anti-double-bubble'' instantons, in which the field is in the false vacuum domain near the poles, and reaches into the domain of the true vacuum on a shell near the equator. It can be cut through the poles to describe shell nucleation, or across the equator, describing the tunneling to true-vacuum inflation in a closed universe.} \\end{figure} \\subsubsection{Open cut} With the analytic continuation used for the previous two instantons, regions I and III will become open inflationary universes in which the field rolls down to the false vacuum. They are separated by the region II, which contains a shell on which $\\phi$ is in the domain of the true vacuum. Therefore we may interpret this solution as the nucleation of a shell of true vacuum on a false vacuum inflationary background, or alternatively, as the creation of such a universe from nothing. Because of the larger action difference, spontaneous shell creation will be quite suppressed compared to bubble formation. \\subsubsection{Closed cut} A more intriguing application of this instanton can be found by choosing a different analytic continuation. Instead of cutting at $\\psi=\\pi/2$, we may choose to leave the three-spheres intact, and cut across the equator. Lorentzian time will be defined by $\\tau=\\tau_{\\rm max} + iT$, and we obtain a metric with closed spacelike sections: \\begin{equation} ds^2 = -dT^2 + a^2(T) d\\Omega_3^2. \\label{eq-closed} \\end{equation} The inflaton field is in the domain of the true vacuum on the nucleation surface (the equator), so it will start rolling down towards the absolute minimum. During this time, the spacelike three-spheres grow exponentially: \\begin{equation} a(T) \\approx H^{-1}(T) \\cosh \\int H(T)\\, dT. \\label{eq-cosh} \\end{equation} Thus we obtain a closed inflationary universe in which the scalar field rolls towards the true vacuum. One could interpret this instanton as describing the creation of such a universe from nothing. But this would just add an alternative to the usual instantons on which the field is entirely in the domain of the true vacuum. A much more interesting interpretation is the one associated with a pre-existing background of false vacuum inflation. In this case, the anti-double-bubble instanton is seen to describe the spontaneous tunneling to the true vacuum in an entire Hubble-volume of de~Sitter space. Unlike the Coleman-De-Luccia bubbles, these regions will not contain an open universe. In the example we are considering, for which Hawking-Moss tunneling is not possible, this shows that one can nevertheless nucleate true vacuum bubbles containing a closed universe. ", "conclusions": "We have described a number of non-singular instantons leading to open inflating universes. They include the Coleman-De-Luccia solution, in which a bubble of true vacuum expands inside a universe inflating in the false vacuum. We found new solutions which contain two bubbles, or a shell of true vacuum. We also constructed instantons with a singularity. If the singularity does not lie on the hypersurface of nucleation, it causes no problems in the Lorentzian region, and can be interpreted as a small region of Planckian density. Such instantons can be used to describe the quantum creation of a closed inflationary universe from space-time foam without the need to use complex solutions. \\subsection*" }, "9803/astro-ph9803107_arXiv.txt": { "abstract": "Using time resolved 2-dimensional aperture photometry we have established that the optical candidate for PSR 1509-58 does not pulse. Our pulsed upper limits ($m_V$ = 24.3 and $m_B$ = 25.7) put severe constraints on this being the optical counterpart. Furthermore the colours of the candidate star are consistent with a main sequence star at a distance of 2-4 kpc. The probability of a chance coincidence with a normal star and the difficulty of explaining the lack of pulsed emission leads us to conclude that this object is an intermediate field star. ", "introduction": "Interest in the optical emission from isolated neutron stars (INSs) has been growing, as recent improvements in detector sensitivity have enabled these faint sources to be observed. Optical observations of neutron stars are important for providing an understanding of pulsar emission mechanisms and allowing direct observations of the energy spectrum of the electron pair plasma. To date 6 INSs have been detected in the optical. Evidence suggests that it is the age and/or the period derivative of the INS, rather than its period, that determines the optical, and indeed multiwavelength, emission from an INS (\\cite{car94a}, \\cite{gol95}). PSR 1509-58 was initially identified by its X-ray emission (\\cite{sew82}) and shortly afterwards a radio signal, with a period of 150ms, was detected (\\cite{man82}). Later studies showed that it had a large $\\dot P$ (1.5 x 10$^{-12}$ ss$^{-1}$ ), in fact the largest known. The pulsar has had its second period derivative measured giving a breaking index ($n=\\omega {\\ddot \\omega}/{\\dot \\omega^2}$ ) of 2.83 $\\pm 0.03$ (\\cite{man85}) in close agreement with the expectations of a radiating dipole (n=3). Its magnetic field ($\\propto (P \\dot P)^{1/2}$) is the largest known and its age 1,600, second only to the Crab pulsar in youth. Its age and location make its association with SNR MSH 15-52 (\\cite{sew84}) at a distance of 4.2 kpc (\\cite{cas75}) likely, although this position has been challenged by \\cite{str94}, who has proposed a greater distance, 5.9 kpc, based upon radio dispersion and x-ray spectra of the extended emission around the pulsar. It has also been observed in soft gamma-rays (\\cite{gun94}) and a tentative optical counterpart has been proposed with $m_V \\approx 22$ (\\cite{car94b}). When taken at a distance of 4.2 kpc the optical observations indicate an absolute magnitude of M$_V$$\\approx 4.9$ (including the effects of interstellar extinction), fainter than the Crab pulsar. However, it is much brighter than would be expected from phenomenological models which have been successful in describing the X-ray emission from PSR 1509-58 (Pacini and Savati 1987). This over-luminosity makes its behaviour similar to the older optical pulsars, PSR0656+14 (\\cite{shear97a}) and Geminga (\\cite{shear97b}). Observations of any pulsed optical component will be crucial in determining what fraction of the emission is thermal and what is magnetospheric. Only the magnetospheric emission would be expected to scale in a manner analogous to that described by Pacini and Savati. Indeed, by considering plausible emission mechanisms (Lu et al 1994) and high energy observations (\\cite{gol95}), it would be reasonable to expect that most of the emission would be non-thermal. Given the importance of a correct determination of the optical emision from PSR1509-58, we have made time-resolved observations in an attempt to confirm its identification and determine the optical pulsed fraction. ", "conclusions": "Our dervived magnitudes and colours, when combined with the lack of optical pulsations, cast doubt on the Caraveo et al (1994b) candidate being the optical counterpart of PSR 1509-58. Our data is consistent with an M type main sequence star at a distance of about 2 kpc. Alternatively, if it is the pulsar then the pulsed fraction must be anomolously low (lower than any other optical pulsar) and most of the radiation thermal. This is contrast with higher energy observations (dominated by non-thermal emission), where the pulsed fraction increases with decreasing energy (\\cite{gre95}). However if the optical emission represents the Rayleigh-Jeans tail of the neutron star's black body spectrum, then with tabulated values for extinction towards MSH15-52, the distance would have to be $\\sim$ 1 kpc assuming a surface temperature of $3~10^6$ K in contrast to the expected distance of at least 4.2 kpc. The presence of a neutron star atmosphere (\\cite{pav96}) would not change this distance sufficiently to explain the optical excess. If the radiation is non-thermal then it is difficult to imagine a geometry and a mechanism which would give such an anomolously low pulsed fraction and still be consistent with high energy observations. Even at the distance derived from radio dispersion (5.9 kpc) the star is still too red to be explained by a thermal extrapolation alone. A final answer to the nature of this star will come from spectroscopy - its magnitude is well within the capabilities of the VLT." }, "9803/astro-ph9803277_arXiv.txt": { "abstract": " ", "introduction": "The observed present-day abundance of rich clusters of galaxies places a strong constraint on cosmology: \\gs$\\Omega^{0.5} \\simeq 0.5$, where \\gs\\ is the {\\em rms} mass fluctuations on 8 \\gh\\ Mpc scale, and \\gW\\ is the present cosmological density parameter (Henry \\& Arnaud 1991, Bahcall \\& Cen 1992, White \\gE\\ 1993, Eke \\gE\\ 1996, Viana \\& Liddle 1996, Pen 1997, Kitayama \\& Suto 1997). This constraint is degenerate in \\gW\\ and \\gs; models with \\gW =1, \\gs \\gi 0.5 are indistinguishable from models with \\gW \\gi 0.25, \\gs \\gi 1. (A \\gs \\gie 1 universe is unbiased, with mass following light on large scales; \\gs \\gie 0.5 implies a mass distribution wider than light). The {\\em evolution} of cluster abundance with redshift, especially for massive clusters, breaks the degeneracy between \\gW\\ and \\gs\\ (see, e.g., Peebles et al. 1989, Oukbir \\& Blanchard 1992, 1997, Eke \\gE\\ 1996, Viana \\& Liddle 1996, Carlberg \\gE\\ 1997, Bahcall \\gE\\ 1997, Fan \\gE\\ 1997, Henry 1997). The evolution of high mass clusters is strong in \\gW\\ =1, low-\\gs\\ (biased) Gaussian models, where only a very low cluster abundance is expected at $z>$0.5. Conversely, the evolution rate in low-\\gW\\ , high-\\gs\\ models is mild and the cluster abundance at $z>$0.5 is much higher than in \\gW=1 models. In Bahcall \\gE\\ (1997) and Fan \\gE\\ (1997) we used the CNOC cluster sample (Carlberg \\gE\\ 1997a,b, Luppino \\& Gioia 1995) to $z \\lesssim$ 0.5 -- 0.8 (with measured masses to $z \\lesssim 0.5$) to decouple \\gW\\ and \\gs: we found \\gW \\gie 0.3 \\gm 0.1 and \\gs \\gie 0.83 \\gm 0.15, consistent with Carlberg \\gE (1997a). The evolution rate, and the distinction among cosmological models, increases with cluster mass and with redshift: in \\gW\\ =1, low-\\gs\\ models, very massive clusters are not expected to exist at high redshifts. In the present paper we extend the previous studies to larger mass and higher redshift clusters, using the three most massive clusters observed so far at high redshifts ($z$ \\gie 0.5--0.9) to independently constrain \\gW\\ and \\gs. The clusters discussed in this paper are the three most massive distant clusters from the EMSS/CNOC sample used above, with masses larger by a factor of $\\sim$ 2 than the mass-threshold used previously (Evrard 1989, Bahcall \\gE\\ 1997, Fan \\gE\\ 1997, Carlberg \\gE\\ 1997a). Reliably measured masses are now available for these clusters from gravitational lensing, temperatures, and velocity dispersions, not previously available in the above studies. Strong Sunyaev-Zel'dovich decrements have also been observed for these clusters, further suggesting that these are massive clusters with large amount of gas. The three clusters have the highest masses (from weak lensing observations), the highest velocity dispersions ($\\sigma_{r} \\gtrsim $1200\\ km $s^{-1}$), and the highest temperatures (T$\\gtrsim $8 kev) in the $z >$ 0.5 EMSS survey (\\S 2). Therefore, they provide a strong constraint on cosmology. We discuss the cluster data in \\S 2 and the cosmological implications in \\S 3. A Hubble constant of $\\rm H_{0}=100\\ h\\ km\\ s^{-1} Mpc^{-1}$ is used. ", "conclusions": "" }, "9803/astro-ph9803195_arXiv.txt": { "abstract": "A measurement of the distance to Virgo Cluster by a direct method along with a realistic error analysis is important for a reliable determination of the value of Hubble Constant. Cepheid variables in the face-on spiral M100 in the Virgo Cluster were observed with the Hubble Space Telescope in 1994 under the HST Key Project on the Extragalactic Distance Scale. This work is a reanalysis of the HST data following our study of the Galactic Cepheids (in an accompanying communication). The periods of the Cepheids are determined using two independent methods and the reasons for varying estimates are analyzed. The log(period) vs $V$-magnitude relation is re-calibrated using LMC data as well as available HST observations for three galaxies and the slope is found to be $-3.45 \\pm 0.15$. A prescription to compute correction for the flux-limited incompleteness of the sample is given and a correction of 0 to 0.28 magnitude in $V$-magnitude for Cepheids in the period range of 35 to 45 days is applied. The extinction correction is carried out using {\\em period vs mean $\\vi $ color} and {\\em $V$-amplitude vs $\\vi $ color at the brightest phase} relations. The distance to M100 is estimated to be $20.4 \\pm 1.7$ (random) $\\pm 2.4$ (systematic) Mpc. ", "introduction": "\\label{sec:intro} A natural scale length for the Universe is provided by the Hubble Constant ($H_0$) and undoubtedly a determination of its value is one of the fundamental problems of cosmology. Over the years, there has been much debate about the value of $H_0$ and the present estimates range from less than 50 \\ksm\\ to over 80 \\ksm. Most probably the major reason for the discrepancy is the conventional distance ladder involving multiple steps. Its main drawback is that analysis of the systematic errors becomes difficult when the calibrating local sample and the observed sample at the next step of the ladder are not identical. Consequently, it is believed that an accurate measurement of the distance to a galaxy cluster which is $\\sim $ 20--30 Mpc away, without involving intermediate steps, will lead to a reliable direct estimate of the value of $H_0$, provided the recession velocity of the cluster is independently known. The Virgo Cluster, which is our nearest cluster of galaxies, is fairly rich in terms of galaxy population, and an average of the distances to the individual galaxies by different methods would provide a good estimate to its mean distance. One of the key projects of the Hubble Space Telescope (HST) was devoted to the calibration of the extragalactic distance scale for a determination of $H_0$ with reasonable accuracy. An examination of the systematic errors in the Cepheid \\plr\\ and measurement of the distance to the Virgo Cluster through Cepheid observation were among the primary aims of this key project. The nearly face-on spiral M100 in the Virgo Cluster was observed on 12 epochs over a span of $\\sim$ 57 days in 1994 with the HST using the filters F555W and F814W, which are almost equivalent to the Johnson V and Cousins I bands respectively (\\cite{freedman:94}). The advantage of choosing this particular galaxy is that being nearly face-on, the errors due to extinction and reddening are expected to be minimal, and further, it is considered to be very similar to the Milky Way in terms of age, chemical composition etc. However, its position relative to the center of the Virgo Cluster is not known accurately, and that introduces some uncertainty in the Virgo distance derived from direct distance estimation to M100. Ferrarese \\ea (1996) reported observations of 70 Cepheids in M100, and obtained a distance of $ 16.1 \\pm 1.3$ Mpc. The value of the Hubble Constant was calculated to be $88 \\pm 24$ \\ksm. On the other hand, Sandage and collaborators re-calibrated the distance to a few galaxies, where supernovae of type Ia were detected earlier, by observing the Cepheids in those galaxies with the HST. They obtained a mean absolute $B$ magnitude at peak of $-19.6$ for normal SN Ia and consequently, a value of $ 52 \\pm 9$ \\ksm\\ for the Hubble Constant (\\cite{sandage:94}; \\cite{saha:94}). However, more recent publications indicate a better reconciliation in the value of $H_0$. Freedman \\ea (1998) summarize a value of $73 \\pm 6$ (statistical) $\\pm 8$ (systematic) \\ksm, as compared to $55 \\pm 3$ (internal) \\ksm\\ quoted by Sandage's group (\\cite{saha:96}). The present work is a re-analysis of the HST data on M100 Cepheids, based on a general calibration of Galactic Cepheids, presented in a companion paper (which we refer henceforth as Paper I). The specific problems we address here are the following: \\bei \\item Period--Luminosity relation applicable to the Cepheids of period $\\ga $ 15 days generally observed in distant galaxies. \\item Importance of the incompleteness correction and quantification of the effect. \\item Uncertainty in the periods of the Cepheids in M100 due to the phase sampling techniques applied as well as the large error in $V$-magnitude, particularly at low flux levels. \\eni The central idea behind distance measurement with Cepheids is the \\plr. However, the values of both the slope and the intercept of this relation continue to be subjects of lively debate. There appears to be a distinct difference in the value of the slope between Cepheids of low and high periods. While applying the \\plr\\ to distant galaxy samples, where only higher period Cepheids can be detected due to flux limitation, this distinction becomes even more crucial. We address this question on the basis of our study of Galactic Cepheids (Paper I) which demonstrates a clear division between two classes of Cepheids, one with periods $\\leq 15$ days, the other at higher periods. The zero-point of the \\plr\\ is another quantity which needs to be fixed unambiguously in order to obtain reliable estimates of distance. We use the recent calibration of the local Cepheids by the Hipparcos mission (\\cite{fc:97}), rather than the distance to the Large Magellanic Cloud which is normally treated as the calibrating point for the distance scale. A crucial aspect of our new analysis is the correction for incompleteness of the Cepheid sample. Since the Cepheid \\plr\\ has an intrinsic scatter due to the finite width of the instability strip at a given period, the Cepheids are observed to be spread over a range of luminosities. All the observed Cepheids in M100 have $V$-magnitudes between 24 and 27 mag. At such faint flux levels it is very likely that for a fixed period, the fainter Cepheids would escape detection, and only the brighter ones will appear in the surveys. This selection bias would have a systematic effect on the period--$V$-magnitude slope, especially at low periods, reminiscent of the Malmquist bias discussed in the literature. In order to counter this effect, one has to take into account the undetected Cepheids, which can be done by adequately correcting the observed magnitudes to fainter levels. Obviously, the amount of correction depends on the scatter of the \\pl\\ diagram. We have devised a formalism to correct for this incompleteness effect which we demonstrate to be present to a large extent in the M100 sample. We have tried to estimate the correction for extinction and reddening, which again, is based on our study of Galactic Cepheids (Paper I). However, in the absence of multi-wavelength observations, this treatment is rather limited, and is based on \\pca\\ relations of Cepheids. Also, for the same reason we could not isolate the extinction correction from the incompleteness correction which ideally we should have been able to. This paper is organized as follows. In Section~\\ref{sec:per_det} we present our methods of determination of Cepheid periods and photometric parameters. The question of choosing the correct \\plr\\ is addressed in Section~\\ref{sec:plr}. In Section~\\ref{sec:incomp}, we devise a formalism for the incompleteness correction of a biased Cepheid sample, and the mathematical aspects of compensation for flux-limited bias are described in the Appendix. Section~\\ref{sec:extcor} deals with the reddening and extinction corrections and the essentials of the numerical methods. The results and major contributions to errors are discussed in Section~\\ref{sec:results} and some remarks on the conclusions are presented in Section~\\ref{sec:summary}. ", "conclusions": "" }, "9803/astro-ph9803126_arXiv.txt": { "abstract": "We revisit the problem of the flat slope of the $Mg_2$ versus $$ relationship found for nuclei of elliptical galaxies (Faber et al. 1992; Worthey et al. 1992; Carollo et al. 1993; Davies et al. 1993), indicating that the Mg/Fe ratio should increase with galactic luminosity and mass. We transform the abundance of Fe, as predicted by classic wind models and alternative models for the chemical evolution of elliptical galaxies, into the metallicity indices $Mg_2$ and $$, by means of the more recent index calibrations and show that none of the current models for the chemical evolution of elliptical galaxies is able to reproduce exactly the observed slope of the $$ versus $Mg_2$ relation, although the existing spread in the data makes this comparison quite difficult. In other words, we can not clearly discriminate between models predicting a decrease (classic wind model) or an increase of such a ratio with galactic mass. The reason for this resides in the fact that the available observations show a large spread due mostly to the errors in the derivation of the $$ index. In our opinion this fact prevents us from drawing any firm conclusion on the behaviour of Mg and Fe in these galaxies. Moreover, as already shown by other authors, one should be careful in deriving trends in the real abundances just from the metallicity indices, since these latter depend also on other physical parameters than the metallicity. This is an important point since abundance ratios have been proven to represent strong constraints for galaxy formation mechanisms. ", "introduction": "Elliptical galaxies do not show the presence of HII regions and it is not possible to resolve single stars in them in order to measure photospheric abundances. Therefore, most of the information on these objects is obtained from their integrated properties: abundances are derived either through colors or integrated spectra and in both cases the derived information is a complicated measure of metallicity and age (the well known age-metallicity degeneracy). The most common metallicity indicators are $Mg_{2}$ and $$, as originally defined in Faber et al. (1977; 1985). Population synthesis techniques are adopted to analyze the integrated properties of ellipticals and to derive an estimate of their real abundances. Unfortunately, they contain several uncertainties residing either in incomplete knowledge of stellar evolution or in deficiencies in stellar libraries, as discussed in Charlot et al. (1996). In recent years more and more population synthesis models (Bruzual and Charlot, 1993; Buzzoni et al., 1992; Bressan et al. , 1994; Gibson, 1997; Bressan et al., 1996; Gibson and Matteucci, 1997; Tantalo et al., 1998) have appeared but the basic uncertainties still remain. In this paper we want to focus our attention about the comparison between theoretical model results and metallicity indicators. In this framework we will analyze the relationship between $$ and $Mg_2$ and its implications for the mechanism of galaxy formation. Several authors (Faber et al. 1992; Worthey et al. 1992; Carollo et al. 1993; Davies et al. 1993; Carollo and Danziger 1994), from comparison of the observed indices with synthetic indices, concluded that the average $[]_{*}$ in giant ellipticals must be larger than the solar value. This result was also confirmed by the analysis of Weiss et al. (1995) who made use, for the first time, of stellar evolutionary tracks calculated under the assumption of non-solar ratios.The same authors found that the $$ versus $Mg_{2}$ relation among nuclei of giant ellipticals is rather flat and flatter than within galaxies. From the flat behavior of $$ vs. $Mg_2$ the same authors inferred that the abundance of Mg should increase faster than the abundance of Fe among nuclei of giant ellipticals. This conclusion is at variance with the predictions of supernova-driven wind models of ellipticals (Arimoto and Yoshii, 1987; Matteucci and Tornamb\\`e 1987). In fact, Matteucci and Tornamb\\`e (1987) showed that, in the framework of the classic wind model for ellipticals, the [Mg/Fe] ratio is a decreasing function of the galactic mass and luminosity. The reason for this behavior is clear: if Fe is mostly produced by the supernovae of type Ia, as it seems to be the case in our Galaxy (Greggio and Renzini 1983a; Matteucci and Greggio 1986), whereas Mg is mostly originating from supernovae of type II, then the iron production is delayed relative to that of Mg and its abundance should be larger in more massive galaxies which develop a wind later than the less massive ones. All of this is valid under the assumption that after the onset of a galactic wind star formation should stop or should be negligible, which is a reasonable assumption for elliptical galaxies. Faber et al. (1992) proposed alternative scenarios to the classic supernova driven wind model, as originally proposed by Larson (1974). They suggested three different scenarios all based on the assumption that Mg is produced by type II supernovae and Fe is mostly produced by type Ia supernovae: i) a selective loss of metals, ii) a variable initial mass function (IMF) and iii) different timescales for star formation. These hypotheses were discussed by Matteucci (1994), who tested them in the context of chemical evolution models. In the hypothesis of the different timescales for star formation Matteucci (1994) suggested that the more massive ellipticals might experience a much stronger and faster star formation than less massive ellipticals leading to a situation where the most massive objects are able to develop galactic winds before the less massive ones. She called this case ``inverse wind model''. On the other hand, in the classic wind model of Larson (1974) the efficiency of star formation was the same for all galaxies thus leading to the fact that the galactic wind in more massive systems occurs later than in less massive ones, due to their deeper potential well. In the models of Arimoto and Yoshii (1987) and Matteucci and Tornamb\\`e (1987) the efficiency of star formation was a decreasing function of galactic mass, based on the assumption that the timescale for star formation is proportional to the cloud-cloud collision timescale which, in turn, is proportional to the gas density. Therefore, since in this monolithic collapse picture the gas density decreases with the galactic mass, the galactic wind was even more delayed for the most massive systems. Matteucci (1994) proposed, as an alternative, a star formation efficiency increasing with the galactic mass and she justified this assumption by imagining that giant elliptical galaxies, instead of forming through a monolithic collapse of a gas cloud, form by merging of gaseous protoclouds. The merging process can, in fact, produce higher densities for increasing galactic mass and/or higher cloud-cloud collision velocities resulting in a faster star formation process. In such a model the galactic wind occurs earlier in massive than in smaller ellipticals thus producing the expected trend of an increasing [Mg/Fe] as a function of galactic mass. Matteucci (1994) also showed that a variable IMF with the slope decreasing with increasing galactic mass and luminosity can produce the same effect without an inverse wind situation. The reason for that resides in the fact that a flatter IMF slope favors massive stars relative to low and intermediate masses, thus favoring Mg production over Fe production. However, Matteucci (1994) could not translate the predicted abundances into $Mg_2$ and $$ since there were no available calibrations for [Fe/H] versus $$ but only calibrations for $[Fe/H]$ vs. $Mg_2$. Therefore she did not compare the predicted abundances with observations. Recently, calibrations for the iron index have become available (Borges et al., 1995; Tantalo et al., 1998) and therefore in this paper we revisit the whole problem of inferring trends on the real abundances by metallicity indices and we discuss the influence of the calibration relationships, which allow us to pass from indices to abundances, and we show that the inferred trend of Mg/Fe with galactic mass is not so clear when interpreted in terms of real abundances, thus warning us from drawing any firm conclusion on galaxy formation processes just on the basis of the observed behavior of $$ versus $Mg_{2}$. indices. The reason for that resides partly in the large spread present in the observational data and partly in the fact that metallicity indices depend not only on the abundances of single elements but also on the ages and on the metallicity distribution (Tantalo et al. 1998) of the different stellar populations present in elliptical galaxies. \\par In Section 2 we will discuss the chemical evolution model, in Section 3 we will define the average abundances of a composite stellar population, in Section 4 we will describe the model results and transform the predicted abundances into indices by means of the most recent metallicity calibrations. Finally in Section 5 some conclusions will be drawn. \\noindent ", "conclusions": "In this paper we have discussed the relation between metallicity indices, such as $Mg_2$ and $$, and total mass in nuclei of ellipticals and its implications in terms of models of formation and evolution of elliptical galaxies. \\par In order to do that we have transformed the average abundance of Fe in the composite stellar population of the galaxy, as predicted by different models of chemical evolution, into $Mg_2$ and $$ indices by means of the available calibrations.\\par We have shown the results of classic wind models for ellipticals, such as those discussed by Arimoto and Yoshii (1987) and Matteucci and Tornamb\\`e (1987), as well as the results of models with variable IMF from galaxy to galaxy and with galactic winds occurring first in the more massive systems, implying that these systems are older than the less massive ones. We have found that it is not possible to establish clearly which kind of model should be preferred, first of all because of the large spread present in the data. Moreover, little difference is found in the predicted indices of models which predict a $[]_{*}$ either increasing or decreasing with galactic mass, although the data seem to suggest an increase of this ratio with galactic mass larger than predicted by any of the models. On this basis, the classic wind model can not be considered worse than the other models. Actually, the classic wind model with a flat constant IMF seems to be the only one which can reproduce the whole range of the observed indices. However, if we isolate the data from Gonzalez (1993) and do not consider the others, then in order to reproduce the flat slope of the $$ versus $Mg_2$ relation, as given by the best-fit of the data, one should assume that Fe among the nuclei of ellipticals is almost constant whereas Mg increases from less massive to more massive nuclei. This is not achieved by any of the models presented here since it would require quite ``ad hoc'' assumptions especially concerning the type Ia SNe. From the numerical experiments performed in this paper we can say that a model which explains at the same time the mass-metallicity and the iron-magnesium relation requires an inverse wind situation, with a strong increase of the star formation efficiency with galactic mass (i.e. Model VI), rather than a variable IMF from galaxy to galaxy, and an IMF with a slope $x=0.8$. However, a model of this type is not able to reproduce the observed ranges of $$ and $Mg_2$. We have also calculated models where amount and concentration of dark matter increases, compatibly with the formulation of the potential energy of the gas, with decreasing galactic luminous mass (see Persic et al. 1996), with the net result of obtaining an ``inverse wind'' situation. The results are very similar to those of Model III. Therefore, to obtain a better agreement with observations one should invoke also in this case an increase of the star formation efficiency with galactic mass. This would certainly flatten the $$ vs. $Mg_2$ relation but it would further shrink the ranges of the predicted indices. In fact, both an increasing star formation efficiency and a decreasing importance of dark matter with increasing luminous galactic mass can be viable solutions to achieve the situation of more massive ellipticals being older than less massive ones. \\par In conclusion, it is quite important to establish the value of [Mg/Fe] from the observational point of view since abundance ratios, such as [Mg/Fe], represent an important diagnostic to infer ages in galaxies, due to the different timescales for the Mg and Fe production. Generally, a high [Mg/Fe] is interpreted as a young age and the upper limit for the age is given by the time at which the chemical enrichment from type Ia SNe starts to become important. This timescale depends not only on the assumed progenitors of type Ia SNe but also on the star formation history of the galaxy considered (see Matteucci 1997) and for giant elliptical galaxies this timescale is of the order of $t_{SNeIa}\\sim 3-4 \\cdot 10^{8}$ years and in any case it can not be larger than 1 Gyr also for smaller systems. This is at variance with what stated by Kodama and Arimoto (1997) who claim, on the basis of results concerning our Galaxy (Yoshii et al. 1996), that this timescale is of the order of 1.5-2.5 Gyr. This is indeed true for our Galaxy where the star formation history has been quite different than in ellipticals and it had been already pointed out in Greggio and Renzini (1983b) and in Matteucci and Greggio (1986). This is a quite important point, both for the galactic chemical enrichment and for the predictions about SN rates at high redshift. Therefore, an enhanced [Mg/Fe] indicates that the process of galaxy formation must have been very fast thus favoring a monolithic collapse scenario rather than a merging scenario. In this framework, a [Mg/Fe] ratio higher in more massive ellipticals than in less massive ones could be interpreted as due to their faster formation and evolution (see Matteucci 1994; Bressan et al. 1996). \\par An independent way of estimating the ages of ellipticals, where for ages we intend the time elapsed from the last star formation event, is to study the $H_{\\beta}$ index. This index is, in fact, related to the age of the dominant stellar population, since it gives a measure of the turn-off color and metallicity. It can therefore be used to solve the age-metallicity degeneracy. Bressan et al. (1996), by analyzing the $H_{\\beta}$ and other physical parameters in the sample of ellipticals observed by Gonzalez (1993), concluded that massive galaxies should have stopped forming stars before less massive ones, in agreement with the results of the inverse wind model discussed here. Finally, we would like to point out that both models with a Salpeter IMF and a variable IMF have a potential problem in reproducing high [$\\alpha$/Fe] ratios in the intracluster medium (ICM), as shown by their low average $[<\\alpha/Fe>]_*$ ratios (see Tables 7-12). Therefore, in agreement with MG95 and Gibson and Matteucci (1997) we conclude that a flat IMF is required to explain the high [$\\alpha$/Fe] ratios, as found by ASCA observations (Mushotzsky 1994)." }, "9803/astro-ph9803256_arXiv.txt": { "abstract": "We introduce and study the distribution of an estimator for the normalized bispectrum of the Cosmic Microwave Background (CMB) anisotropy. We use it to construct a goodness of fit statistic to test the coadded 53 and 90 GHz {\\it COBE}-DMR 4 year maps for non-Gaussianity. Our results indicate that Gaussianity is ruled out at the confidence level in excess of 98$\\%$. This value is a lower bound, given all the investigated systematics. The dominant non-Gaussian contribution is found near the multipole of order $\\ell=16$. Our attempts to explain this effect as caused by the diffuse foreground emission from the Galaxy have failed. We conclude that unless there exists a microwave foreground emission which spatially correlates neither with the DIRBE nor Haslam maps, the cosmological CMB anisotropy is genuinely non-Gaussian. ", "introduction": "We shall consider fluctuations in the CMB as a random field on the sphere, $\\frac{\\Delta T}{T}({\\bf n})$. One can expand such a field in terms of Spherical Harmonic functions: \\begin{eqnarray} \\frac{\\Delta T}{T}({\\bf n})=\\sum_{\\ell m}a_{\\ell m}Y_{\\ell m}({\\bf n}) \\label{almdef} \\end{eqnarray} For a statistically isotropic field one has \\begin{eqnarray} \\langle a_{\\ell_1 m_1}a^*_{\\ell_2 m_2}\\rangle=C_{\\ell_1} \\delta_{\\ell_1 \\ell_2} \\delta_{m_1 m_2} \\label{defiso} \\end{eqnarray} We can also define the two-point function in terms of $\\frac{\\Delta T}{T}({\\bf n})$. Isotropy implies that the correlation matrix can only depend on the angle between the two points considered. This is encoded in the 2-point correlation $C^{(2)}(\\theta)$. From (\\ref{almdef}) and (\\ref{defiso}) we find \\begin{equation}\\label{c2cl} C^{(2)}(\\theta)={\\sum _\\ell}{2\\ell+1\\over 4\\pi}C_\\ell P_\\ell(\\cos\\theta) \\end{equation} Hence the $C_\\ell$ may be regarded as a Legendre transform of the 2-point correlation function. It is a standard lore that, barring some mathematical obstructions, one can reconstruct the probability distribution function of any random field from its moments. Isotropy imposes ``selection rules'' on these moments. For instance, the 3-point moment is given by \\begin{eqnarray} \\langle a_{\\ell_1 m_1}a_{\\ell_2 m_2} a_{\\ell_3 m_3}\\rangle= \\left ( \\begin{array}{ccc} \\ell_1 & \\ell_2 & \\ell_3 \\\\ m_1 & m_2 & m_3 \\end{array} \\right ) C_{\\ell_1\\ell_2\\ell_3} \\label{defnpoint} \\end{eqnarray} where the $(\\ldots)$ is the Wigner $3J$ symbol. The coefficients $C_{\\ell_1\\ell_2\\ell_3}$ are usually called the bispectrum. If we assume that there are no correlations between different $\\ell$ multipoles then the only non-zero component of the bispectrum is $C_{\\ell\\ell\\ell}=B_\\ell$. The collapsed 3-point correlation function $C^{(3)}(\\theta)$ (the average of a temperature squared at one point, and a temperature at another point, separated by an angle $\\theta$) is now \\begin{eqnarray} C^{(3)}(\\theta)={\\sum _\\ell}{\\left(2\\ell+1\\over 4\\pi\\right)}^{3/2} \\left ( \\begin{array}{ccc} \\ell & \\ell & \\ell \\\\ 0 & 0 & 0 \\end{array} \\right )B_\\ell P_\\ell(\\cos\\theta) \\end{eqnarray} in analogy with (\\ref{c2cl}). Hence the $B_\\ell$ is related to the Legendre transform of $C^{(3)}$. The angular power spectrum $C_\\ell$ is often considered a more powerful tool than the correlation function $C^{(2)}(\\theta)$ for discriminating between theories, and one might argue the same way with regard to the reduced bispectrum $B_{\\ell}$ and the 3-point function $C^3(\\theta)$. The importance of higher order statistics for characterizing large scale structure has been stressed before (\\cite{lss}). The non-linear evolution of primordial Gaussian fluctuations has been analysed in detail (\\cite{lss,bouch92}) and the skewness arising in such models has been shown to be consistent with current observations (\\cite{bouch93,gaz94}). \\cite{luo94} discussed the statistical properties and detectability of the bispectrum for a variety of non-Gaussian signals. \\cite{kog96a} measured the pseudocollapsed and equilateral three point function of the DMR four year data and found them to be consistent with Gaussianity. The analysis performed here should be considered complementary to that of \\cite{kog96a}: non-Gaussian signals which may be obscured in real space can become evident in $\\ell$ space. In this letter we shall use a general formalism for generating estimators of higher order moments on a sphere (\\cite{fergormag}). In this formalism one considers all possible tensor products of $\\Delta T_\\ell$ (each multipole component of the field) and from these one extracts the singlet (invariant) term. In the case of bispectrum one has \\begin{eqnarray} {\\hat B}_\\ell&=&\\alpha_\\ell\\sum_{m_1m_2m_3}\\left ( \\begin{array}{ccc} \\ell & \\ell & \\ell \\\\ m_1 & m_2 & m_3 \\end{array} \\right ) a_{\\ell m_1}a_{\\ell m_2} a_{\\ell m_3} \\nonumber \\\\ \\alpha_\\ell&=&\\frac{1}{(2\\ell+1)^{\\frac{3}{2}}}\\left ( \\begin{array}{ccc} \\ell & \\ell & \\ell \\\\ 0 & 0 & 0 \\end{array} \\right )^{-1} \\label{bispec} \\end{eqnarray} Note that only even values of $\\ell$ lead to nonzero values of the ${\\hat B}_\\ell$ due to the symmetries of the Wigner 3-J coefficients. In practice it is essential to factor out the power spectrum from our statistic. We also wish to define statistics which are invariant under parity transformations, and not just rotations. Therefore we define $I^3_\\ell$ to be \\begin{eqnarray} I^3_\\ell &=&\\left| { {\\hat B}_\\ell\\over ({\\hat C}_\\ell)^{3/2}} \\right| \\label{defI} \\end{eqnarray} where ${\\hat C}_\\ell=\\frac{1}{2\\ell+1}\\sum_m|a_{\\ell m}|^2$. Our statistics are dimensionless and are normalized so that a cylindrically symmetric multipole has $I^3_\\ell=1$. The $\\ell=2$ case was discussed and given a physical interpretation in \\cite{mag1}. The quadrupole has 5 degrees of freedom. Of these only 2 are rotationally invariant. One is the quadrupole intensity $C_2$, and tells us how much power there is in the quadrupole. The other is essentially $I^3_2$ and tells us how this power is distributed among the different $a_{2 m}$ but only as far as there is a rotationally invariant meaning to the concept. For instance if $I^3_2=1$ then there is a frame in which all the power is concentrated in the $m=0$ mode. Such a quadrupole is cylindrically symmetric, but of course the symmetry axis orientation is uniformly distributed, to comply with statistical isotropy. If $I^3_2=0$ then on the contrary cylindrical symmetry is maximally broken. The probability distribution function of $I^3_2$ is uniform in Gaussian theories (\\cite{mag1}). ", "conclusions": "The result that we have obtained raises a number of questions which we shall attempt to answer. From Fig.~\\ref{fig1} it is clear that $I^3_{16}$ is far in the tail of the Gaussian ensemble and it dominates the statistic. One would like to understand the importance of both cosmic variance and noise to this measurement. We would also like to assess the extent to which a galactic foreground contaminant could be responsible for this result. In order to answer the first question we look for Bayesian estimates for the $I^3_\\ell$ as they are {\\it for our sky}. To do this we first estimate what the temperature fluctuations $T_i=\\frac{\\Delta T}{T}({\\bf n}_i)$ in each pixel $i$ in our dataset are likely to be, given DMR observations $O_i$, and noises $\\sigma_i^2$. We construct the posterior $P(T_i|O_i)$ assuming uniform priors in $T_i$, and also that a priori no correlations exist between the $T_i$. The latter assumption is often used in image restoration algorithms, such as maximum entropy methods. We then produce an ensemble of skies with the distribution $P(T_i|O_i)$. From it we infer $P(I^3_\\ell|O_i)$, the distributions for what the $I^3_\\ell$ for our sky are likely to be given DMR observations and noise. This procedure will allow us to assess the importance of noise in each of our measurements. However note that this analysis is totally decoupled from the result in the previous section where all we need to know are the observed $I^3_\\ell$, not their estimates for our sky. In Fig.~\\ref{fig2} we plot in dotted lines $P(I^3_\\ell|O_i)$ for our data set. We also plot in solid lines the cosmic variance distribution of $I^3_\\ell$ in skies with the same galactic cut. The vertical line is the observed invariant $I^3_\\ell(O_i)$. As expected we see that, as $\\ell$ gets larger, the spread in $P(I^3_\\ell|O_i)$ due to noise becomes more important, at $\\ell=18$ dominating the distribution function. On the other hand we clearly have succeeded in making measurements for $\\ell=4,6,8,12,14,16$. For them $P(I^3_\\ell|O_i)$ are peaked and clearly different from the cosmic variance distribution. The fact that $P(I^3_{16}|O_i)$ does not peak at $I^3_{16}(O_i)$ is merely a failure of the prior. The measurement of $I^3_{16}(O_i)$ is therefore a signal and is not dominated by noise. We have further checked that the signal to noise in power at $\\ell=16$ is of order 1. Next we wish to know if galactic emissions could be blamed for this result. We can proceed in three ways. Firstly we may use instead the DMR cosmic emission maps, where a linear combination of the various DMR channels is used to separate out the foreground Galactic contamination. In these maps the noise level is considerably higher. Plotting the counterpart of Fig.~\\ref{fig2} for this case we find that the distributions of the actual $I^3_\\ell$ for our sky, given noise induced errors, are very similar to their cosmic variance distributions. The measurement is therefore dominated by noise and inconclusive. We find $X^2_{COBE}=.4$, consistent with Gaussianity, but this is a mere check of the Gaussianity of noise. Hence this approach towards foregrounds turns into a dead end, but serves to show how large angle Gaussian tests is a field constrained by noise, not cosmic variance. As an alternative approach we may subject galactic templates to the same analysis. At the observing frequencies the obvious contaminant should be foreground dust emission. The DIRBE maps (\\cite{boggess92}) supply us with a useful template on which we can measure the $I^3_\\ell$s. We have done this for two of the lowest frequency maps, the $100$ $\\mu$m and the $240$ $\\mu$m maps. The estimate is performed in exactly the same way as for the DMR data (i.e. using the extended Galaxy cut). We performed a similar exercise with the Haslam 408Mhz (\\cite{haslam}) map. We display their values in Fig.~\\ref{fig3}. As expected the two maps have consistent values for the $I^3_\\ell$. However they do not have a non-Gaussian value at $\\ell=16$. Indeed for all $\\ell$ the $I^3_\\ell$ are within Gaussian cosmic variance error bars. This is not surprising. DIRBE maps exhibit structures on very small scales. These should average into a Gaussian field when subject to a $7^\\circ$ beam. As a third alternative we may use foreground corrected maps. In these one corrects the coadded 53 and 90 Ghz maps for the DIRBE correlated emission. We have considered corrected maps in ecliptic and galactic frames, and also another map made in the ecliptic frame but with the DIRBE correction forced to have the same coupling as determined in the galactic frame. As shown in Fig.~\\ref{fig3}, in all of these the non-Gaussian signal at $\\ell=16$ is enhanced, although we observe large variations in $I^3_\\ell$ at $\\ell=4-8$ (a phenomenon noticed before when estimating $C_\\ell$-s). In fact the corrected maps exclude Gaussianity at the confidence level of 99.5\\%. It would be interesting to relate our result to the curious dip in power at $\\ell\\approx 16$ provided by the maximum likelihood estimates in \\cite{gorski97}. These show that, {\\it assuming a Gaussian signal}, the power in signal and noise is unusually low at $\\ell\\approx 16$. One wonders how this would be affected if non-Gaussian degrees of freedom were allowed into the estimation (\\cite{fergormag}). We have also subjected our work to a variety of numerical tests. Arbitrary rotations of the coordinate system affects results to less than a part in $10^5$. More importantly, comparing data pixelized in the ecliptic and galactic frames, we found that our results were very robust, indeed more so than the power spectrum estimation (see the bottom pannel of Fig.~\\ref{fig3}). We also tried different galactic cuts, and found that although the non-Gaussian signal gets transferred to other $\\ell$, one does not fully erase it until a cut of $\\pm 40^\\circ$ is applied. Finally we checked the effect of varying the offset in the cut map. We found that for any other prescription than the one used the effect is enhanced, often leading to rejecting Gaussianity at more than the 99.5\\% confidence level. To conclude, we have not been able to attribute our result to a known contaminating source or a systematic. Indeed the confidence level quoted refers to the worst result obtained within the set of effects explored. Of course it is always possible that this non-Gaussian signal comes from some yet unmapped foreground, which cannot be separated from the CMB anisotropy signal in the {\\it COBE}-DMR data --- the poorly known free-free emission from the Galaxy comes to mind here. If indeed our results are due to a foreground contamination one should note the following two points. First, we would have demonstrated that DMR data is more contaminated by foregrounds than thought before. Second, Galactic emissions on the scales considered are often assumed to be Gaussian. In fact this assumption is used in subtraction algorithms based on the idea of optimal filtering. The discovery of a distinctly non-Gaussian galactic emission would in the very least require a rethinking of the foreground subtraction algorithms. If, on the other hand, the CMB signal itself is demonstrably non-Gaussian, we would not need to over-emphasise the epistemological implications of our findings." }, "9803/astro-ph9803074_arXiv.txt": { "abstract": "\\noindent A phase transition in the nature of matter in the core of a neutron star, such as quark deconfinement or Bose condensation, can cause the spontaneous spin-up of a solitary millisecond pulsar. The spin-up epoch for our model lasts for $2\\times 10^7$ years or 1/50 of the spin-down time (Glendenning, Pei and Weber in Ref. \\cite{glen97:a}). The possibility exists also for future measurements on X-ray neutron stars with low-mass companions for mapping out the tell-tale ``backbending'' behavior of the moment of inertia. Properties of phase transitions in substances such as neutron star matter, which have more than one conserved charge, are reviewed. ", "introduction": "Neutron stars have a high enough interior density as to make phase transitions in the nature of nuclear matter a distinct possibility. Examples are hyperonization, negative Bose condensation (like $\\pi^-$ and $K^-$) and quark deconfinement. According to the QCD property of asymptotic freedom, the most plausible is the quark deconfinement transition. From lattice QCD simulations, this phase transition is expected to occur in very hot ($T\\sim 200$ MeV) or cold but dense matter. In this work we will use the deconfinement transition as an example, but in principle, any transition that is accompanied by a sufficient softening of the \\eos and occurs at or near the limiting mass star, can produce a similar signal. The paper is organized as follows. We discuss first the physical reason why a rapidly rotating pulsar, as it slows down over millions of years because of angular momentum loss through the weak electromagnetic process of magnetic dipole radiation, will change in density due to weakening centrifugal forces and possibly encounter, first at its center, and then in a slowly expanding region, the conditions for a phase transition. Conversely, an accreting star will be spun up from low to high frequency by accretion from a low-mass companion. This too will have a very long time-scale because accretion is regulated by the radiation pressure of the star's surface, heated by infalling matter. After having discussed the reasons why we might see signals of phase changes, both in rapidly rotating stars that are spinning down because of angular momentum loss to radiation and stars that are spinning up due to the input of angular momentum by accretion, we discuss some aspects of phase transitions that are common to all first order transitions in neutron star matter, or more generally in isospin asymmetric matter. ", "conclusions": "" }, "9803/astro-ph9803242_arXiv.txt": { "abstract": "We have developed 1D time-dependent numerical models of accretion discs, using an adaptive grid technique and an implicit numerical scheme, in which the disc size is allowed to vary with time. The code fully resolves the cooling and heating fronts propagating in the disc. We show that models in which the radius of the outer edge of the disc is fixed produce incorrect results, from which probably incorrect conclusions about the viscosity law have been inferred. In particular we show that outside-in outbursts are possible when a standard bimodal behaviour of the Shakura-Sunyaev viscosity parameter $\\alpha$ is used. We also discuss to what extent insufficient grid resolutions have limited the predictive power of previous models. We find that the global properties (magnitudes, etc. ...) of transient discs can be addressed by codes using a high, but reasonable, number of fixed grid points. However, the study of the detailed physical properties of the transition fronts generally requires resolutions which are out of reach of fixed grid codes. It appears that most time-dependent models of accretion discs published in the literature have been limited by resolution effects, improper outer boundary conditions, or both. ", "introduction": "The thermal-viscous accretion-disc instability model is more than 15 years old (see Cannizzo \\shortcite{can93b} for a historical overview). It is widely accepted that it provides the correct description of dwarf-nova outbursts and of (`soft') X-ray transient events. When, however, observations of these systems are compared with predictions of the model, the agreement is far from perfect (e.g. Lasota \\& Hameury \\shortcite{lh98} and references therein). It is sometimes also unclear what the predictions of the model are. One of the reasons for these uncertainties is the existence of various, different, versions of the model. From the very beginning these versions of the disc instability model differed in assumptions about viscosity and boundary conditions; they differed in the amount of matter accreted during the outburst, the shapes of light-curves, etc. (see Cannizzo \\shortcite{can93b}). At that time these differences seemed to be less important than the differences between the disc instability model and the competing, mass-transfer instability model \\cite{bp81}. The exponentially decaying tails of theoretical light curves predicted by the mass-transfer model were thought to contradict observations \\cite{can93b} and the outer disc radius behaviour during and after outbursts seemed to favour the disc instability model \\cite{io92}. The demise of the mass-transfer instability model was, however, caused by the lack of a physical mechanism which would trigger it. With one model left it became important to establish just what its predictions are, and not merely whether it is better (or worse) than the competing model \\cite{pvw86}. The first systematic study of the disc instability model was presented by Cannizzo \\shortcite{can93a}, who analysed the importance of various terms in the disc evolution equations and the influence of the numerical grid resolution on the outburst properties. Ludwig, Meyer-Hofmeister \\& Ritter \\shortcite{lmr94} studied general properties of disc outbursts, such as the location of the instability that triggers them. Recently Ludwig \\& Meyer \\shortcite{lm98} analysed non-Keplerian effects which may arise during front propagation. The general conclusions of this group of studies were that non-Keplerian effects are negligible, that a few hundred grid points provide a sufficient resolution for the calculation results to be independent of the number of grid points, and, finally, that with the usual assumption (see Smak (1984b) of a jump in the value of the viscosity parameter $\\alpha$, the model produces only `inside-out' outbursts, i.e. outbursts starting in the inner disc regions. This last conclusion, if true, would entail changing the standard viscosity law (in which the $\\alpha$ parameter is constant in the hot and cool branch of the $\\Sigma$ - $T_{\\rm eff}$ curve) because outburts starting in the outer disc regions are clearly observed in classical dwarf-nova system SS Cyg \\cite{mau96}. This law already had to be modified when it was found \\cite{sma84b} that in order to get lightcurves similar to those observed in dwarf novae, $\\alpha$ in outburst had to be larger than $\\alpha$ in quiescence. The absence of `outside-in' outbursts in these studies, however, is just the result of keeping the outer disc radius constant in the calculations (Section 4.1) \\cite{io92,sma84b}; from this point of view, there is no reason to modify the viscosity prescription. This does not in itself prove that changes in viscosity are correctly described by the bimodal behaviour of the $\\alpha$-parameter (see e.g. Gammie \\& Menou 1998), but the reasons given for preferring other versions (the exponential decay from outburst being the principal one) are not compelling, and these versions involve more fundamental changes in the disc physics (see Lasota \\& Hameury 1998 for discussion and references). For example, Cannizzo, Chen \\& Livio \\shortcite{ccl95} use the formula $\\alpha = \\alpha_0 \\left(H/R \\right)^n$, but to make the model work they have to `switch off' convection. It is interesting, therefore, to recall that Faulkner, Lin \\& Papaloizou (1983) found dwarf nova outburst with $\\alpha$ constant, but their model was criticized \\cite{can93b} because they claimed that convection has only a minor influence on the energy transport in the disc. These are not formal problems because the constant $\\alpha$ models predict optically thin quiescent discs, whereas in bimodal $\\alpha$ models the quiescent disc is optically thick. There is observational evidence that dwarf nova discs in quiescence are optically thin (see Horne, 1993 and references therein). Conclusions about the number of grid points required to get resolution-invariant results seem, on inspection, too optimistic, especially because fronts are not resolved, a point which is particularily worrying for the heating fronts. The present situation of the disc instability model seems to be confused. Various versions are based on different assumptions about the physical processes in the disc and numerical codes suffer either from incorrect boundary conditions or from insufficient resolution or from both. Quite often, in the case of explicit codes the resolution is limited by the required computer time. In this article we describe a numerical model of time-dependent accretion discs, using an adaptive grid technique and an implicit numerical scheme, in which the disc size is allowed to vary with time. This numerical scheme allows rapid calculations of disc outburst cycles at very high resolution. These properties allow an easy comparison with other versions of the model and a systematic study of its various assumptions. In the near future we will use our code to model various properties of dwarf novae and X-ray transients. The model was alread used to model properties and outbursts of the dwarf-nova WZ Sge \\cite{lhh,hlh} and the rise to outburst of the X-ray transient GRO J1655-40 \\cite{hlmn97}. In \\S 2 we discuss the time-dependent equations describing the disc radial structure and the implicit method used to solve them with a high spatial resolution. The vertical structure of the disc, and hence the heating and cooling terms that enter the time-dependent energy equation, are considered in \\S 3. In \\S 4 we present the results of our calculations and we discuss the importance of having sufficient numerical resolution and a correct boundary condition at the outer edge of the disc. ", "conclusions": "We have constructed a numerical code which can calculate, in a reasonable amount of computer time and at very high spatial resolutions, long cycles of accretion disc outbursts. This code works efficiently in the most general framework of the disc instability model and does not require special assumptions about viscosity or outer or inner radii. Its validity is of course limited by the way physical processes such as turbulent viscosity, convection, radiative transfer etc. are treated. Of course it is also a 1D code modeling a fundamentally 2D (or even 3D) situation. Since the mass of the central object enters the disc equations only in $\\Omega_K$, we expect most of our results to be valid for BH disc models as well ({\\i.e. similar}, but at a slightly smaller radius), as long as irradiation and general relativistic effects can be neglected. In future work we intend to include effects of irradiation (Dubus et al. 1998) and to apply the code to a systematic study of dwarf nova outbursts and X-ray transient events. \\subsection*" }, "9803/astro-ph9803132_arXiv.txt": { "abstract": "We have studied the poor southern cluster of galaxies S639. Based on new Str\\\"{o}mgren photometry of stars in the direction of the cluster we confirm that the galactic extinction affecting the cluster is large. We find the extinction in Johnson B to be $\\AB = 0.75\\pm 0.03$. We have obtained new photometry in Gunn $r$ for E and S0 galaxies in the cluster. If the Fundamental Plane is used for determination of the relative distance and the peculiar velocity of the cluster we find a distance, in velocity units, of $(5706\\pm 350)\\kms$, and a substantial peculiar velocity, $(839\\pm 350)\\kms$. However, the colors and the absorption line indices of the E and S0 galaxies indicate that the stellar populations in these galaxies are different from those in similar galaxies in the two rich clusters Coma and HydraI. This difference may severely affect the distance determination and the derived peculiar velocity. The data are consistent with a non-significant peculiar velocity for S639 and the galaxies in the cluster being on average 0.2 dex younger than similar galaxies in Coma and HydraI. The results for S639 caution that some large peculiar velocities may be spurious and caused by unusual stellar populations. ", "introduction": "The relation known as the Fundamental Plane (FP) may be used for determination of relative distances to E and S0 galaxies (e.g., Dressler et al.\\ 1987; J\\o rgensen, Franx \\& Kj\\ae rgaard 1996, hereafter JFK96; Baggley 1996; Hudson et al.\\ 1997). The FP relates the effective radius, $\\re$, the mean surface brightness within this radius, $\\Ie$ and the (central) velocity dispersion $\\sigma$, in a relation, which is linear in logarithmic space (Djorgovski \\& Davis 1987; Dressler et al.\\ 1987). The FP has a low scatter ($15-20$\\% in $\\re$) and is therefore a valuable tool for studies of peculiar velocities of galaxies and clusters (e.g., Baggley 1996; Hudson et al.\\ 1997). The use of the FP for determination of distances and peculiar velocities relies on the assumption that the FP is universally valid. Several authors have investigated possible differences in the FP related to the cluster environment (e.g., Burstein, Faber \\& Dressler 1990; Lucey et al.\\ 1991; de Carvalho \\& Djorgovski 1992; JFK96; Baggley 1996). Only de Carvalho \\& Djorgovski find that the environment has significant effects on the FP. These authors find field galaxies to be brighter than cluster galaxies of similar effective radii and velocity dispersions. de Carvalho \\& Djorgovski also find field galaxies to be bluer and have weaker $\\Mgtwo$ line indices than cluster galaxies with similar velocity dispersions. This is in general agreement with studies that show that E and S0 galaxies in the outer parts of clusters have weaker $\\Mgtwo$ and $\\Fe$ indices than those in the central parts of clusters (Guzm\\'{a}n et al.\\ 1992; JFK96; J\\o rgensen 1997). In this paper we study the poor cluster of galaxies S639, previously studied by JFK96. The cluster identification is from Abell, Corwin \\& Olowin (1989). The cluster has a radial velocity in the Cosmic Microwave Background (CMB) frame of $cz_{\\rm CMB}=6545\\kms$ and is located $\\approx 28$$\\degr$ from the direction to the large mass-concentration known as the ``Great Attractor'' (Faber \\& Burstein 1988). The velocity dispersion of the cluster is $456_{-74}^{+83}\\kms$ (JFK96). Its richness is 14 measured as the number of galaxies with magnitudes between $m_3$ and $m_3+2$ (Abell et al.\\ 1989). $m_3$ is the magnitude of the third ranked galaxy. S639 has a smaller velocity dispersion and is poorer than clusters like the Coma cluster and the HydraI cluster. Coma and HydraI have velocity dispersions of $1010_{-44}^{+51}\\kms$ and $608_{-39}^{+58}\\kms$, respectively (Zabludoff, Huchra \\& Geller 1990). The richnesses given by Abell et al.\\ (1989) is 106 for Coma and 39 for HydraI. Using the FP for 10 E and S0 galaxies in S639 JFK96 found a large peculiar velocity of the cluster, $v_{\\rm pec}=(1295 \\pm 359)\\kms$ relative to the CMB frame. Further, JFK96 found that the galaxies in the cluster follow a $\\Mgtwo$-$\\sigma$ relation offset from the relation established for their full sample of 11 clusters. The galaxies in S639 had on average weaker $\\Mgtwo$ indices, see also J\\o rgensen (1997). JFK96 tried to correct the derived peculiar velocity of the cluster for the offset in the $\\Mgtwo$ indices by including a $\\Mgtwo$ term in the FP. The result was $v_{\\rm pec}=(879\\pm 392)\\kms$. However, the coefficient for the $\\Mgtwo$ term is not well determined, cf.\\ JFK96. S639 is located at low galactic latitude, ($l$,$b$) = ($280\\degr$,$11\\degr$). Thus, the galactic extinction is large and uncertainties in the adopted value may severely affect the precision of the derived distance and peculiar velocity for the cluster. The main issue discussed in this paper is whether the large peculiar velocity of S639 found by JFK96 is real or the result was caused either by incorrect correction for the (large) galactic extinction, by selection effects, or by unusual stellar populations. In order to reach conclusions about these issues we have obtained additional photometry of galaxies in S639, giving a sample of 21 E and S0 galaxies with available photometric and spectroscopic parameters. We have also obtained Str\\\"{o}mgren $uvby$-$\\beta$ photometry for stars in the direction of the cluster. This photometry is used to determine the galactic extinction affecting the cluster. The sample selection for the E and S0 galaxies and the available data are briefly described in Sect.\\ 2. The determination of the galactic extinction is covered in Sect.\\ 3. In Sect.\\ 4 the FP is discussed and used for determining the distance to the cluster. The importance of the stellar populations is investigated in Sect.\\ 5. The conclusions are summarized in Sect.\\ 6. The relations between the parameters for the galaxies established in this paper are determined by minimization of the sum of the absolute residuals perpendicular to the relations. This fitting technique has the advantage that it is rather insensitive to a few outliers, and that it treats the coordinates in a symmetric way. The uncertainties of the coefficients are derived by a bootstrap method. See also JFK96 for a discussion of this fitting technique. ", "conclusions": "The galactic extinction in the direction of the poor cluster of galaxies S639 has been determined from Str\\\"{o}mgren $uvby$-$\\beta$ photometry for stars in the direction of the cluster. Further, we have tested the consistency of the derived galactic extinction by using the $(B-r)$-$\\Mgtwo$ relation for E and S0 galaxies in the cluster. Our best estimate of the galactic extinction in the direction of the cluster is $\\AB = 0.75\\pm 0.03$. The FP for S639 has been established based on a sample of 21 E and S0 galaxies. The coefficients for the FP for this cluster are not significantly different from those of the FP for the Coma and the HydraI clusters, and are also in agreement with previous results for other nearby clusters (e.g., JFK96). Under the assumption that the FP (coefficients and zero point) is universal we find a distance, in velocity units, to S639 of $(5706\\pm 350) \\kms$. This implies a peculiar velocity for the cluster of $(839\\pm 350)\\kms$. The E and S0 galaxies in S639 have significantly smaller $\\Mgtwo$ indices, larger $\\HbG$ indices and are bluer than E and S0 galaxies of similar velocity dispersions in Coma and HydraI. The offset in the FP for S639 relative to Coma and HydraI may be due to a difference in the stellar populations, rather than a large peculiar velocity for S639. The data are consistent with a zero peculiar velocity and mean ages of the S639 galaxies 0.2 dex younger than the mean ages of similar galaxies in Coma and HydraI. Alternatively, the $\\Mgtwo$ indices and the $(B-r)$ colors are consistent with a metallicity difference of 0.1 dex, with galaxies in S639 having a lower metallicity than those in Coma and HydraI. In this case the peculiar velocity of S639 is $\\approx 490\\kms$. However, this interpretation is not consistent with the strong $\\HbG$ indices measured for the four galaxies, for which we have measurements of this index. We conclude that the peculiar velocity of S639 is most likely overestimated if the FP is used as a distance determinator for this cluster. Even though many studies have shown that the FP (and the $\\Dn$-$\\sigma$ relation which is a projection of the FP) to a large degree is universally valid (e.g., Burstein, Faber \\& Dressler 1990; JFK96; Baggley 1996), our results for S639 caution that there may be exceptions (see also Gregg 1992). When distance determinations are attempted the best approach will be to obtain colors and line indices together with the other required data. This will give the possibility of identifying clusters (and galaxies), which deviate strongly from the mean relations between the various global parameters. These clusters can also be expected to deviate from the FP otherwise valid for the bulk of the E and S0 galaxies. One may attempt to include a $\\Mgtwo$ term in the FP in order to correct for the effects caused by differences in the stellar populations. However, the coefficient for such a term is not well determined, cf.\\ JFK96, and the peculiar velocities derived with this method may not be accurate enough for investigations of large-scale flows. \\vspace{0.5cm} Acknowledgements: Lars Freyhammer is thanked for obtaining part of the observations used for this research. The Danish Board for Astronomical Research and the European Southern Observatory are acknowledged for assigning observing time for this project and for financial support. Support for this work was provided by NASA through grant number HF-01073.01.94A to IJ from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. HJS acknowledges financial support from the Carlsberg Foundation, Denmark." }, "9803/astro-ph9803304_arXiv.txt": { "abstract": "\\B observed several galactic binary X--ray pulsars during the Science Verification Phase and in the first year of the regular program. The complex emission spectra of these sources are an ideal target for the \\B instrumentation, that can measure the emission spectra in an unprecedented broad energy band.\\\\ Using this capability of \\B a detailed observational work can be done on the galactic X--ray pulsars. In particular the 0.1--200 keV energy band allows the shape of the continuum emission to be tightly constrained. A better determination of the underlying continuum allows an easier detection of features superimposed onto it, both at low energy (Fe K and L, Ne lines) and at high energies (cyclotron features).\\\\ We report on the spectral properties of a sample of X--ray pulsars observed with \\B comparing the obtained results.\\\\ Some ideas of common properties are also discussed and compared with our present understanding of the emission mechanisms and processes. ", "introduction": "The instrumentation aboard \\B \\cite{sax,lecs,mecs,hp,pds} is particularly well suited to study the X--ray emission from X--ray pulsars. This class of sources is composed by binary systems in which a magnetized rotating neutron star accretes matter from a less evolved mass--donor star. The mass--donor may be a OB supergiant star as in the case of Vela X--1, a Be main--sequence or near--main--sequence star as in the case of transient recurrent pulsars like A0535+26, a low mass star as in the case of 4U1626--67. The type of mass donor star strongly affects the temporal behaviour on the medium (days) to long (years) time scales. The transient behaviour is almost completely restricted to the subclass of X--ray pulsars that have Oe or Be counterparts. \\B has observed some persistent pulsars and one transient pulsar during the first year of its operative life. We report results from the observations of some of these sources, emphasizing the commonalities and the differences. In particular we discuss the observational evidence on cyclotron line feature, comparing the observed results, also in terms of possible correlations, with the expected ones on the basis of the available theoretical models. ", "conclusions": "The correlation showed in Figure 1 was ``qualitatively'' predicted by M\\'esz\\'aros and Nagel \\cite{mn} (see also \\cite{pulmod}). The model predicts a width of the cyclotron feature proportional to its energy and to the square root of the electron temperature of the atmosphere \\begin{equation} \\label{eq:fw} \\Delta \\omega_B \\simeq \\omega_B \\left( 8 \\times \\ln(2) \\times \\frac{\\rm kT_e}{\\rm m_ec^2} \\right)^{\\frac{1}{2}} |\\cos\\theta| \\label{eq13} \\end{equation} In this equation $\\Delta\\omega_B$ is the line width, $\\omega_B$ is the cyclotron line frequency, $T_e$ is the electron temperature and $\\theta$ is the viewing angle with respect to the magnetic field axis. A better insight on the properties of the cyclotron lines can be obtained with pulse--phase resolved spectroscopy, as equation \\ref{eq:fw} suggests that there may be a dependence on the viewing angle of the observed line width (see also \\cite{pphspe}). However Araya and Harding (1996) \\cite{araya} caution that, in the limit of a single scattering, the line width is not related to the electron temperature. This ambiguity in the interpretation of these observational data points out the need of a more detailed and quantitative model for the line properties and for the broad band continuum emission of X--ray pulsars" }, "9803/astro-ph9803148_arXiv.txt": { "abstract": "The Goddard High Resolution Spectrograph (GHRS) of the {\\it Hubble Space Telescope (HST)} has been used to observe the boron 2500 \\AA\\ region of \\bd-13. At a metallicity of [Fe/H]=$-3.00$ this is the most metal-poor star ever observed for B. Nearly 26 hours of exposure time resulted in a detection. Spectrum synthesis using the latest Kurucz model atmospheres yields an LTE boron abundance of log \\eps(B)$= +0.01\\pm0.20$. This value is consistent with the linear relation of slope $\\sim$1.0 between log \\eps(B$_{\\rm LTE}$) and [Fe/H] found for 10 halo and disk stars by Duncan \\etal\\ (1997). Using the NLTE correction of Kiselman \\& Carlsson (1996), the NLTE boron abundance is log \\eps(B)$= +0.93\\pm0.20$. This is also consistent with the NLTE relation determined by Duncan \\etal\\ (1997) where the slope of log \\eps(B$_{\\rm NLTE}$) vs. [Fe/H] is $\\sim$0.7. These data support a model in which most production of B and Be comes from the spallation of energetic C and O nuclei onto protons and He nuclei, probably in the vicinity of massive supernovae in star-forming regions, rather than the spallation of cosmic ray protons and alpha particles onto CNO nuclei in the general interstellar medium. ", "introduction": "The light elements lithium, beryllium, and boron are of great interest out of proportion to their very low abundances, having implications in Big Bang Nucleosynthesis and stellar structure, as well as in constraints on models of galactic chemical evolution. The ``canonical'' theory of the origin of the elements Li, Be, and B was first presented by Reeves, Fowler, \\& Hoyle (1970) and further developed by Meneguzzi, Audouze, \\& Reeves (1971), and then Reeves, Audouze, Fowler, \\& Schramm (1973). In this model, most light element formation can be accounted for by galactic cosmic rays (GCR) impinging on the interstellar medium (ISM), assuming a constant flux of GCRs through the life of the Galaxy and making reasonable assumptions about CR confinement by the Galactic magnetic field. Meneguzzi \\etal\\ (1971) also introduced the idea of a large (up to three orders of magnitude) increase in the low energy (5-40 MeV nucleon$^{-1}$) CR flux; since CRs in this energy range are mostly shielded from the Solar System by the solar wind, they are not detectable. This additional CR flux increased the production of all light elements and matched the isotopic ratios and total abundances to the accuracy known at the time. Reeves \\& Meyer (1978) added the additional constraint that models should match not only present-day abundances but their evolution throughout the life of the Galaxy. Their conclusions were similar to MAR, except that they had to introduce infall of light-element-free matter into the Galactic disk to match the evolution with time. In retrospect, it can be seen that the data they were fitting were sparse and not very precise. With the launch of the the {\\it Hubble Space Telescope} ({\\it HST}) and the availability of uv-sensitive CCD detectors (B is usually observed at $\\lambda$2500 and Be at $\\lambda$3130), data are now much more numerous and accurate, and abundances can be traced from the epoch of formation of the Galactic halo until the present day. In the past several years, the evolution of Li, Be, and B has been used as a test of different models of the chemical and dynamical evolution of the Galaxy. For example, in the models of Vangioni-Flam \\etal\\ (1990), Ryan \\etal\\ (1992), and Prantzos \\etal\\ (PCV; 1993), light element production depends on the intensity and shape of the GCR spectrum, which in turn depends on the supernova (SN) and massive star formation rates. It also depends on the rise of the (progenitor) CNO abundances and the decline of the gas mass fraction, which is affected by rates of infall of fresh (unprocessed) material and outflow, e.g. by SN heating. Other things being equal, at early times when target CNO abundances were low, light element production would be much lower for a given CR flux than presently, when the ISM abundances are higher. PCV found that even with these numerous adjustable parameters, no time-independent CR spectrum can reproduce the evolution of light element abundances. By assuming a very particular form of time variation of the CR flux (greatly enhanced at early epochs), they were able to (barely) fit the evolution of the abundances. The present investigation supports a different solution to the problem of the origin of the light elements. Duncan \\etal\\ (1997) present B abundances in a large number of stars ranging in metallicity from $\\sim$solar to [Fe/H]~$\\sim -2.8$. They find that (LTE) B follows metals in direct proportion from the earliest times (very metal-poor stars) to the present, with little if any change of slope between halo and disk metallicities. A straightforward interpretation of this is that the rate of production of B and Be does {\\it not} depend on the CNO abundances in the ISM, and that the production site is associated with the production site for metals. This would be true if the spallation process most important for light element production is not primarily protons and $\\alpha$ particles colliding with CNO nuclei in the ISM but rather C and O nuclei colliding with ambient protons and $\\alpha$ particles, probably in regions of massive star formation (cf. Vangioni-Flam \\etal\\ 1996 and Ramaty \\etal\\ 1997). This paper focuses specifically on the B measurement in \\bd-13; it is consistent with (and was used to help determine) the relationships between LTE and NLTE B and [Fe/H] seen in Duncan \\etal\\ (1997). The data suggest that B production at the lowest metallicities occurs in the same way as today, and thus support the new description of galactic light element production. It is possible that recent GRO satellite observations of gamma rays from the Orion Nebula (Bloemen \\etal\\ 1994; Bykov \\& Bloemen 1994) provide direct evidence of C and O spallation occurring today, even though not all instruments on GRO detected evidence of such spallation (Murphy \\etal\\ 1996). The light element data alone, however, are the strongest evidence in support of a new model of their production. ", "conclusions": "Fig.~4 shows the LTE abundances as a function of [Fe/H] from stars analyzed in the larger investigation of Duncan \\etal\\ (1997), with the point for \\bd-13 from the present investigation emphasized. It can be seen that there is an approximately linear relation between $\\log$ \\eps(B$_{\\rm LTE}$) and [Fe/H] over both disk and halo metallicities, and that \\bd-13\\ is consistent with this relationship. A least-squares fit to all the data of Fig.~4 (allowing for errors in both coordinates) yields a slope of 0.96$\\pm$0.07 and a reduced chi square, $\\chi_{\\nu}^2$, of 0.71, indicating an excellent fit. If NLTE abundances are used, as shown in Fig.~5, the slope is 0.70$\\pm$0.07 and $\\chi_{\\nu}^2 = 1.63$. Although it is true that \\bd-13\\ is used to determine this line, it is important to note that this star is quite consistent with the trend defined by the other stars. Inspection and $\\chi_{\\nu}^2$ tests confirm this. \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics{figs/fig4.eps}} \\caption{LTE B abundances from Duncan \\etal\\ (1997) with the program star highlighted.} \\end{figure} \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics{figs/fig5.eps}} \\caption{NLTE B abundances from Duncan \\etal\\ (1997) with the program star highlighted.} \\end{figure} \\subsection{Comparison to standard models and a new model for light element production} The slope of close to 1 suggesting a primary process is {\\it not} expected from canonical models of CR spallation in the ISM, which predict a secondary process and thus a steeper relation. In a secondary process the rate of light element production depends on the product of the abundance of target CNO nuclei and the CR flux, both of which vary with time. If SNe are the source of the target nuclei and the ISM is well-mixed, the ISM metallicity is proportional to the integral (total) number of SN up to a given time. If, as is commonly supposed, SNe also seed the acceleration mechanism which produces CRs, the CR flux is proportional to the SN rate. The result is light element abundances which vary quadratically with the metallicity of the ISM, or a logarithmic slope of 2 (Prantzos \\etal\\ 1993). Figs.~4 and 5 show that such a slope is certainly not consistent with our data. Duncan \\etal\\ (1997) discuss these issues in greater detail. As was discussed by Duncan, Lambert, and Lemke (1992), the data for the first three metal-poor stars observed for B already seemed to show a linear (in the log) relationship with [Fe/H], suggesting some primary production mechanism rather than the secondary mechanism in the ISM described above. This idea has been modelled in detail by Cass\\'e \\etal\\ (1995), Ramaty \\etal\\ (1995 and 1997), Lemoine \\etal\\ (1997), and Vangioni-Flam \\etal\\ (1996). In the new scenario, B and Be are primarily produced by the spallation of C and O onto protons and $\\alpha$ particles. Such a process could occur near massive star SNe, where the particle flux would be very non-solar in composition; depleted in H and He and especially enriched in O and C. Vangioni-Flam \\etal\\ find that a composition matching either winds from massive (Wolf-Rayet; WR) stars in star-forming regions or massive star SNe produce a flux of O and C which, after further acceleration, can reproduce both the magnitude and slope of B production seen in Figs.~4 and 5 through collisions with protons and $\\alpha$ particles. As Ramaty \\etal\\ point out, production of some additional $^{11}$B by the neutrino process (Woosley \\etal\\ 1990) is not ruled out, and may be favored on energetic grounds. Nevertheless the bulk of the B and Be would be produced from the spallation process. Although the NLTE correction to the B abundance of \\bd-13 is relatively large and tends to raise the B abundance above the line in Fig.~5, two other effects not included here would tend to move it closer to the curve. One is the effect of the blending Co line discussed above, which could reduce the B abundance as much as $\\sim$0.15 dex. The other is the fact that if the spallation producing light elements is caused by O (and to a lesser extent C), $\\log$ \\eps(B) should be plotted against [O/H] rather than [Fe/H]. As is well-known, very metal-poor stars are overabundant in O compared to Fe (moving points representing the most metal-poor stars to the right in a figure with O on the x-axis). Duncan \\etal\\ (1997) make such a plot, and demonstrate that although the measurement errors in O are greater than those for Fe, when all the metal-poor stars are considered together a straight line of slope 1.10$\\pm$0.14 fits the LTE abundances, and one of slope 0.82$\\pm$0.10 the NLTE abundances. However, oxygen abundance measurements are also surrounded by greater uncertainty than are iron measurements, and systematic errors in the oxygen abundance which depend on metallicity will affect the derived slope." }, "9803/astro-ph9803181_arXiv.txt": { "abstract": "Following an approach developed by Paczy\\'nski \\& Stanek, we derive a distance to the Large Magellanic Cloud (LMC) by comparing red clump stars from the {\\em Hipparcos}\\/ catalog with the red clump stars observed in two fields in the LMC that were selected from the ongoing photometric survey of the Magellanic Clouds to lie in low extinction regions. The use of red clump stars allows a single step determination of the distance modulus to the LMC, $\\mu_{0,LMC} = 18.065\\pm 0.031\\pm 0.09\\;$mag (statistical plus systematic error), and the corresponding distance, $R_{LMC}= 41.02\\pm 0.59\\pm 1.74\\;kpc$. This measurement is in excellent agreement with the recent determination by Udalski et al., also based on the red clump stars, but is $\\sim 0.4\\;$mag smaller than the generally accepted value of $\\mu_{0,LMC} = 18.50\\pm 0.15\\;$mag. We discuss possible reasons for this discrepancy and how it can be resolved. ", "introduction": "The generally accepted distance modulus to the Large Magellanic Cloud (LMC) is $\\mu_{0,LMC} \\approx 18.5 \\pm0.15\\;$mag (for recent discussion see Westerlund 1997, Madore \\& Freedman 1998). However, there is a long standing $\\sim 0.3\\;$mag discrepancy between the ``long'' distance determined using Cepheids (e.g. Laney \\& Stobie 1994) and the ``short'' distance determined using RR Lyr stars (e.g. Walker 1992, Layden et al.~1996). A similar discrepancy is present in the distance to the LMC derived with the supernova SN1987A ($\\mu_{0,LMC}< 18.37\\;$mag, Gould \\& Uza 1998; $\\mu_{0,LMC} = 18.56\\;$mag, Panagia et al.~1997). Recently Udalski et al.~(1998) used red clump stars observed in the LMC by the OGLE 2 project (Udalski et al.~1997) and obtained a value of $\\mu_{0,LMC} = 18.08\\pm 0.03 \\pm 0.12 \\;$mag (statistical plus systematic error). This distance modulus is $\\sim 0.4\\;$mag smaller than the ``long'' distance modulus used, for example, by the {\\em HST}\\/ Extragalactic Distance Scale Key Project team (e.g.~Rawson et al.~1997 and references therein). Because errors in the distance to the LMC can propagate into errors in such key quantities as distances, luminosities, masses, and sizes of extragalactic objects, it is important to check the result of Udalski et al.~(1998) using independent data, in order to investigate possible systematic errors. Red clump stars are the metal rich equivalent of the better known horizontal branch stars, and theoretical models predict that their absolute luminosity only weakly depends on their age and chemical composition (Seidel, Demarque, \\& Weinberg 1987; Castellani, Chieffi, \\& Straniero 1992; Jimenez, Flynn, \\& Kotoneva 1998). Indeed, the absolute magnitude-color diagram from {\\em Hipparcos}\\/ data (Perryman et al.~1997, their Figure~3) clearly shows a compact red clump -- the variance in the $I$-band magnitude is only $\\sim 0.15\\;$mag (Stanek \\& Garnavich 1998; Udalski et al.~1998). Despite their large number and the theoretical understanding of their evolution, red clump stars have seldom been used as distance indicators. However, Stanek (1995) and Stanek et al.~(1994, 1997) used these stars to map the Galactic bar. Paczy\\'nski \\& Stanek (1998) used the red clump stars observed by the OGLE project (Udalski et al.~1993) to obtain the distance to the Galactic center. Stanek \\& Garnavich (1998) used red clump stars observed by the {\\em HST}\\/ in M31 to obtain a one-step distance to this galaxy. In this paper we follow the approach of Paczy\\'nski \\& Stanek (1998) and present an estimate of the distance to the LMC based on the comparison between the red clump giants observed locally by the {\\em Hipparcos}\\/ (Perryman et al.~1997) satellite and those observed in the LMC by the $UBVI$ digital photometric survey of the Magellanic Clouds (Zaritsky et al.~1997). In Section 2 we describe the data used in this paper and select low extinction regions for further analysis. In Section 3 we analyze the red clump distribution in the LMC and derive the distance to this galaxy. In Section 4 we discuss the possible reasons for the discrepancy with the Cepheid distance to the LMC and how it can be resolved. ", "conclusions": "As with all distance-ladder techniques, our analysis includes the assumption that the calibrating and target objects being compared are intrinsically similar. In our red clump analysis, this assumption is manifested by the assertion that the $I$-band brightness of red clump stars is independent of the age, chemical composition, and mass differences that may exist between the red clump stars near the Sun and those in the LMC. Indeed, the LMC red clump is systematically bluer than the local one, indicating the somewhat different properties of these stars. However, Paczy\\'nski \\& Stanek (1998), Stanek \\& Garnavich (1998) and Udalski et al.~(1998) found that the $I$-band peak magnitude of the red clump depends very weakly on their $(V-I)_0$ color in the range $0.7<(V-I)_0<1.4$, and therefore is independent of the metallicity (Jimenez et al.~1998). This is confirmed in this paper as well, by comparing the peak brightness of the red clump for two color ranges $0.55<(V-I)_0 <0.8$ and $0.8<(V-I)_0 <1.25$ (see the previous Section). The fact that the observed red clump distributions are so narrow ($\\sigma_{RC}\\approx 0.15\\;$mag) indicates that the age dependence of the red clump $I$-band peak luminosity is also small ($\\lesssim 0.1\\;$mag). Otherwise, in a system with a complex star formation history, such as the LMC (Holtzman et al.~1997; Geha et al.~1998), the resulting red clump should have considerable width. Stanek \\& Garnavich (1998) compared three different lines-of-sight that probe a large range of M31 galactocentric distances and locations, and hence a range of metallicities and possibly ages and star formation histories. The fact that the derived distance moduli for their three fields varied by only $\\sim 0.035\\;$mag indicates that the red clump is a potentially stable standard candle. The mostly empirical support for using the red clump stars as a distance indicator should also be verified using modern theory of the stellar structure and evolution. In particular, $I$-band predictions are seldom given by such theoretical calculations. So why does the red clump distance to the LMC disagree with the Cepheid distance (Madore \\& Freedman 1998)? As usual, there are several possible answers. Contrary to our arguments given above, there might still be something ``unusual'' about the red clump population in the LMC. Although the red clump and Cepheid distances to the LMC are discrepant, the distances to M31 derived from the two methods are in excellent agreement ($m-M = 24.471\\pm0.035\\pm0.045$ from Stanek \\& Garnavich~1998 and 24.44$\\pm 0.13$ from Freedman \\& Madore~1990). Another possibility is that the Cepheid distance to the LMC is simply poorly determined, as there are few Cepheids with well determined parallaxes in the {\\em Hipparcos}\\/ catalog. In their recent study, Madore \\& Freedman (1998) find $\\mu_{0,LMC}= 18.44\\pm 0.35\\;$mag, from a sample of 19 Cepheids observed by {\\em Hipparcos}\\/ with good $BV$ data, and $\\mu_{0,LMC}= 18.57\\pm 0.11\\;$mag, from a sample of only 7 Cepheids with good $BVIJHK$ data. Yet a third possibility, as discussed by Madore \\& Freedman (1998), is that there are other effects on the Cepheid PL relation (e.g. extinction, metallicity, and statistical errors), which are as significant as any reassessment of the zero point based on {\\em Hipparcos}. The metallicity effect on the Cepheid PL relation, determined by Kennicutt et al.~(1998) ($\\delta(m - M)_0/\\delta[O/H] = -0.24 \\pm 0.16 $\\magdex), reduces the discrepancy between the red clump and Cepheid distances to the LMC by $\\sim 0.1\\;$mag, while the somewhat larger metallicity dependence found by Sasselov et al.~(1997) and Kochanek (1997) reduces it by $\\sim 0.15\\;$mag. To illustrate the effect of the assumed reddening on the derived distance modulus, we note that the value of the LMC distance modulus, $\\mu_{0,LMC}= 18.54\\;$mag, derived recently by Salaris \\& Cassisi (1998) and based on the $V$-band brightness of the RR Lyr stars, becomes $\\mu_{0,LMC}= 18.22\\;$mag if their assumed reddening of $E(B-V)=0.10\\;$mag is increased to $E(B-V)=0.20\\;$mag, corresponding to the mean reddening found by Harris et al.~(1997). It is disturbing that the distance to a key calibrator of the entire distance scale is uncertain by up to 20\\%. As described by Udalski et al.~(1998), the $\\sim 0.4\\;$mag discrepancy between their (and now our as well) ``short'' distance to the LMC and the ``long'' distance to the LMC from the Cepheids can be resolved by using detached eclipsing binaries as a direct distance indicator (Paczy\\'nski 1997). The Cepheids in the LMC can also be used to get a direct distance estimate through a modified Baade-Wesselink method (e.g. Krockenberger 1996; Krockenberger, Sasselov, \\& Noyes 1997). Both these methods require no intermediate steps in the distance ladder, therefore avoiding the propagation of errors usually crippling the distance scale. With the 6.5--8 meter telescopes now being built in the Southern Hemisphere the necessary spectroscopy of the detached eclipsing binaries and Cepheids can be quite easily obtained for these 14--18 magnitude stars. It is worth mentioning here that the effort to obtain direct distances with the detached eclipsing binaries and Cepheids to the M31 and M33 galaxies is already under way and the first results look promising (project DIRECT: Kaluzny et al.~1998, Stanek et al.~1998, Krockenberger et al.~1998, Sasselov et al.~1998). To summarize, among the various stellar distance indicators the red clump giants might be the best for determining the distance to the LMC and other nearby galaxies because there are so many red clump stars. In particular, {\\em Hipparcos}\\/ provided accurate distance determinations for almost 2,000 such stars, but unfortunately $I$-band photometry is available for only $\\sim 30\\%$ of them, so it is important to obtain $I$-band photometry for all {\\em Hipparcos}\\/ red clump giants. We also need to test the metallicity dependence of {\\em Hipparcos}\\/ red clump giant absolute luminosities and colors. There are many stars within $100\\;pc$ of the Sun for which very high-resolution spectroscopy is possible." }, "9803/astro-ph9803238_arXiv.txt": { "abstract": "Recent work by Pringle and by Maloney, Begelman \\& Pringle has shown that geometrically thin, optically thick, accretion disks are unstable to warping driven by radiation torque from the central source. This work was confined to isothermal (\\ie surface density $\\Sigma\\propto R^{-3/2}$) disks. In this paper we generalize the study of radiation-driven warping to include general power-law surface density distributions, $\\Sigma\\propto R^{-\\delta}$. We consider the range $\\delta=3/2$ (the isothermal case) to $\\delta=-3/2$, which corresponds to a radiation-pressure-supported disk; this spans the range of surface density distributions likely to be found in real astrophysical disks. In all cases there are an infinite number of zero-crossing solutions (\\ie solutions that cross the equator), which are the physically relevant modes if the outer boundary of the disk is required to lie in a specified plane. However, unlike the isothermal disk, which is the degenerate case, the frequency eigenvalues for $\\delta\\neq 3/2$ are all distinct. In all cases the location of the zero moves outward from the steady-state (pure precession) value with increasing growth rate; thus there is a critical minimum size for unstable disks. Modes with zeros at smaller radii are damped. The critical radius and the steady-state precession rate depend only weakly on $\\delta$. An additional analytic solution has been found for $\\delta=1$. The case $\\delta=1$ divides the solutions into two qualitatively different regimes. For $\\delta \\ge 1$, the fastest-growing modes have maximum warp amplitude, $\\beta_{\\rm max}$, close to the disk outer edge, and the ratio of $\\beta_{\\rm max}$ to the warp amplitude at the disk inner edge, $\\beta_o$, is $\\gg1$. For $\\delta < 1$, $\\beta_{\\rm max}/\\beta_o\\simeq 1$, and the warp maximum steadily approaches the origin as $\\delta$ decreases. This implies that nonlinear effects {\\it must} be important if the warp extends to the disk inner edge for $\\delta \\ge 1$, but for $\\delta < 1$ nonlinearity will be important only if the warp amplitude is large at the origin. Because of this qualitative difference in the shapes of the warps, the effects of shadowing of the central source by the warp will also be very different in the two regimes of $\\delta.$ This has important implications for radiation-driven warping in X-ray binaries, for which the value of $\\delta$ characterizing the disk is likely to be less than unity. In real accretion disks the outer boundary condition is likely to be different from the zero-crossing condition that we have assumed. In accretion disks around massive black holes in active galactic nuclei, the disk will probably become optically thin before the outer disk boundary is reached, while in X-ray binaries, there will be an outer disk region (outside the circularization radius) in which the inflow velocity is zero but angular momentum is still transported. We show that in both these cases the solutions are similar to the zero-crossing eigenfunctions. ", "introduction": "Evidence for warped, precessing accretion disks in astrophysical systems ranging from X-ray binaries to active galactic nuclei has steadily accumulated over the last two decades (see Maloney \\& Begelman 1997a, and references therein). The origin and maintainence of such warped disks has until recently stood as an unsolved theoretical problem. While it is possible, for example, to generate non-planar modes with $m=1$ symmetry in thin, relativistic disks (\\cite{k90}; \\cite{kh91}), these modes only exist at small radii ($R\\lessapprox 10$ Schwarzschild radii), since they rely on trapping of the modes in the non-Newtonian region of the potential. However, an important clue was provided by \\cite{pet77}, who pointed out that in an optically thick disk with a central source of luminosity, the pressure resulting from re-radiation of the intercepted flux will produce a net torque if the disk is warped. Almost twenty years were to pass before it was recognized that radiation pressure torque actually leads to a warping instability. \\cite{pri96} (P96) showed that, for the special case in which the disk surface density $\\Sigma\\propto R^{-3/2}$ (corresponding to an isothermal disk in the usual $\\alpha-$disk formalism, with disk viscosity $\\nu\\propto R^{3/2}$), even an initially planar disk is unstable to warping by this mechanism. Pringle solved the linearized twisted disk equations in this case using a WKB approximation. Maloney, Begelman \\& Pringle (1996, Paper I, hereafter MBP) found exact solutions to the linearized twist equations, and demonstrated the importance of the outer boundary condition for determining the growth rates of the unstable modes. These previous works all specialized to the case of an isothermal disk, which simplifies the twist equations. While this may be a reasonable approximation for some astrophysical disks (\\eg the masing molecular disk in NGC 4258; see MBP), there are many other systems, such as accretion disks in X-ray binary systems, where this is likely to be a poor assumption. In this paper we extend the work of MBP by considering disks with power-law surface density profiles, $\\Sigma\\propto R^{-\\delta}$. We consider the range $-3/2\\le\\delta\\le 3/2$: the lower limit corresponds to a radiation-pressure-supported disk (\\eg \\cite{fkr92}), while the upper limit is the isothermal value (MBP). Within the limitations of assuming a constant power-law for the surface density, this spans the probable range of surface density laws relevant to real astrophysical accretion disks. For example, the standard Shakura-Sunyaev gas pressure-supported disk is characterized by $\\delta=0.75$ (\\cite{ss73}). In \\S 2 we discuss the twist equation, including the effect of radiation torque, and cast it into a more convenient form. We solve the equation numerically in \\S 3 and discuss both the time-dependent and steady-state solutions. As in the isothermal case, the outer boundary condition is crucial for determining the stability of the disk and the growth rates of the unstable modes. In \\S 4 we discuss the important issue of the appropriate outer boundary condition for accretion disks around stellar-mass objects and AGN. Finally, in \\S 5 we discuss the implications of the results and their application to real accretion disks. ", "conclusions": "Earlier work on the radiation-driven warping instability discovered by Pringle (P96; MBP) considered only the isothermal, $\\delta=3/2$ case. In this paper we have considered more general power-law disk density distributions, from the isothermal disk to $\\delta=-3/2$, corresponding to a radiation-pressure supported disk; this spans the range that is likely to be relevant to astrophysical disks. Although the shapes of the eigenfunctions do change with decreasing $\\delta$, the most important features of the instability are generic. Most importantly, the instability exists over the entire range of surface density index that we have considered, and the critical radius above which disks are unstable to radiation-driven warping changes only by a factor of $\\simeq 6$ from $\\delta=3/2$ to $\\delta=-3/2$. Similarly, the growth and precession rates (in dimensional units) do not depend strongly on $\\delta$ (see the discussion after equation [18] and below). Evaluating equation (15) for the critical radius, \\bea R_{\\rm cr}&=&\\left(5.9\\times10^8\\;\\,\\quad -\\quad 3.5\\times 10^9\\right)\\; \\left({\\eta\\over \\epsilon_{0.1}} \\right)^2\\left({M\\over\\msol}\\right)\\;{\\rm cm} , \\nonumber \\\\ &=& \\left(5.9\\times10^{16}\\quad - \\quad 3.5\\times 10^{17}\\right)\\, \\left({\\eta\\over \\epsilon_{0.1}} \\right)^2\\left({M\\over 10^8 \\msol}\\right)\\;{\\rm cm} , \\eea where the range in numerical values is for $\\delta=3/2$ to $\\delta=-3/2$ and $\\epsilon=0.1\\epsilon_{0.1}$. The only warping modes with zeros at $R < R_{\\rm cr}$ are damped, so that disks that are smaller than $R_{\\rm cr}$ are stable against warping. In consequence of the $\\epsilon^{-2}$ scaling of $R_{\\rm cr}$, accretion disks in systems with very low radiative efficiency will not be unstable to radiation-driven warping unless they are implausibly large. For this reason, this mechanism cannot provide an explanation for the warp in the thin maser disk of NGC 4258 (\\eg \\cite{miy95}, \\cite{her97}) if the inner disk is advection-dominated with $\\epsilon\\sim 10^{-3}$ (\\cite{las96}; see the discussion in MBP), since the maser disk would be far too small for instability in this case. This also indicates that radiation-driven warping generally will not be important in cataclysmic variables or protostellar disks dominated by accretion-powered luminosity, since the radiative efficiency is limited to small values as the stellar surfaces are at $R_*\\gg R_s$ (but see \\cite{arm97} for a discussion of the possible action of the instability in the protostellar case). To evaluate the typical precession timescales, we need to evaluate the viscous inflow timescale at $R_{\\rm cr}$. Letting $\\nu_1=\\alpha c_s H$, where $c_s$ is the isothermal sound speed and $H$ is the scale height, we can write the viscous timescale as \\be t_{\\rm visc}\\sim {2\\over 3} {R\\over V_\\phi}\\alpha^{-1} (H/R)^{-2} \\ee where $V_\\phi$ is the rotational velocity (assumed to be Keplerian) and $H/R$ is evaluated at the radius in question. Since $t_{\\rm visc}\\propto R^{3/2}$, and $R_{\\rm cr}/R_o=x_{\\rm cr}^2$, \\be t_{\\rm visc}(R_{\\rm cr})\\sim {1\\over 3}\\left({\\eta\\over\\epsilon} \\right)^3 {R_s\\over\\alpha c} x_{\\rm cr}^{3/2}\\left(H/R\\right)^{-2} \\ee where $H/R$ is now evaluated at $R_{\\rm cr}$. Taking the precession timescale $t_{\\rm prec}=2\\pi/\\sigma_r$, where $\\sigma_r$ is given by equation (18), and evaluating the constants, we find \\be t_{\\rm prec}\\sim 12\\, {\\eta^2\\over\\epsilon_{0.1}^3} {M/\\msol\\over \\alpha_{0.1}}\\left({H/R\\over 0.01}\\right)^{-2}_{R_{\\rm cr}}\\;{\\rm days} \\ee with only weak dependence on $\\delta$: the numerical coefficient only varies by a factor of two over the whole range of $\\delta$. Thus the precession timescales for X-ray binary systems (the only systems in which precession can actually be observed) are expected to be of the order of weeks to months. This is of course the precession timescale for the steady-state modes from linear theory. As discussed in \\S 3, real disks will ordinarily be unwarped beyond some maximum radius, either the physical edge of the disk or where the disk becomes optically thin. This outer boundary, which will not in general correspond to the critical radius, will determine the warp growth rate. We expect that the warp will eventually saturate at some amplitude (but see Pringle 1997). Assuming that the disk does reach a steady state, what will the precession rate be? There is reason to suspect it may not be very different from the linear theory result. Figure 6 shows that, except for growth rates very close to the maximum, the real part of the eigenvalue $\\tilde\\sigma_r$, \\ie the precession rate, is nearly independent of the growth rate. In the isothermal case, in fact, $\\tilde\\sigma_r$ is independent of $\\tilde\\sigma_i$. This suggests that, however different modes may couple in reaching the final state, the precession rate will be similar to the linear steady-state result. Implicit throughout this paper has been the assumption that the disks are optically thick to both absorption and re-emission, so that they are subject to the radiation-driven warping instability. This requirement imposes a minimum mass accretion rate that must be exceeded for the disk to be optically thick. In Appendix D, we derive this critical mass accretion rate for three different possible sources of opacity in astrophysical disks (electron scattering, dust absorption, and Kramer's opacity) and show that it does not in general place any significant limitations on occurrence of the instability. As discussed in \\S 3.2, there is one very important systematic change in the nature of the instability with $\\delta$. The difference in the behavior of the growing modes for $\\delta \\ge 1$ and $\\delta < 1$ is of fundamental importance for the evolution of disks warped by radiation pressure. For $\\delta \\ge 1$, the fast-growing modes all have their maximum warp (\\ie tilt $\\beta$) close to the outer edge of the disk, and the amplitude $\\beta_{\\rm max}$ is much greater than $\\beta_o$, the tilt at the origin. This immediately implies that the warp must reach the nonlinear regime when the tilt at small radius is negligible. In this case the evolution of the disk at radii interior to the warp maximum is almost certainly driven by the nonlinear evolution of the outer warp (\\eg \\cite{pri97}), so that nonlinear effects {\\it must} be important if the warp extends to the disk inner edge. For $\\delta < 1$, the behavior is qualitatively different, as $\\beta_{\\rm max}/\\beta_o$ is always of order unity. In this regime, nonlinearity will be important only if the warp has grown out of the linear regime at the origin. Furthermore, because the shapes of the growing warps in these two regimes are so dissimilar, the effects of shadowing of the central source by the warping of the disk will be very different. These distinctions are liable to be crucial for X-ray binaries such as SS 433 and Her X-1, which show evidence for a {\\it global} precessing warp. One final point regarding X-ray binary systems must be mentioned. In one of the best-studied systems, Her X-1, the direction of precession of the warp is inferred to be retrograde with respect to the direction of rotation (\\eg \\cite{ger76}) and this has also been suggested for SS 433 (\\cite{leib84}; \\cite{bri89}). As shown in Appendix B, in the absence of external torques the direction of precession of the warp must be prograde. However, the qualification on this statement is extremely important: as pointed out in \\S 4.2, and discussed in detail by Maloney \\& Begelman (1997b), including the quadrupole torque from a companion star allows retrograde as well as prograde solutions to exist. The zero-crossing outer boundary condition that we have imposed will not be strictly correct in real astrophysical disks. However, as discussed in \\S 4, the solutions that obey the likely realistic outer boundary conditions -- the optically thin outer boundary for accretion disks in active galactic nuclei, and a flat outer boundary for disks in X-ray binaries -- are in all important respects similar to the zero-crossing solutions. Radiation-driven warping and precession offers a robust mechanism for producing tilted, precessing accretion disks, in accreting binary systems such as Her X-1 and SS 433, and in active galactic nuclei. Because radiation-driven warping is an inherently global mechanism, it avoids the difficulties inherent in other proposed mechanisms for producing warping and precession, \\eg communicating a single precession frequency through a fluid, differentially-rotating disk. This mechanism can thus explain the simultaneous precession of inner disks (as evidenced by the jets of SS 433 and the pulse profile variations of Her X-1) and outer disks (as required to match the periodicities in X-ray flux and disk emission in these same objects). A full understanding of the nature of the radiation-driven warping instability will require nonlinear simulations of the type presented in Pringle (1997), which will not only allow for inclusion of the nonlinear terms but also inherently nonlinear effects such as shadowing. This will be the subject of future work." }, "9803/astro-ph9803328_arXiv.txt": { "abstract": " ", "introduction": "A reconstruction of density perturbation spectrum is a key problem of the modern cosmology. It made a dramatic turn after detecting the primordial CMB anisotropy by DMR COBE (Smoot et al\\cite{1}, Bennet et al\\cite{2}) as the signal found at $10^0$ $\\Delta T/T = 1.06 \\times 10^{-5}$ appeared to be few times more than the expectable value of $\\Delta T/T$ in the most simple and developed cosmological model -- standard CDM one (SCDM\\footnote {$\\Omega_M = 1$, $\\Omega_b = 0.06$ (Walker at al\\cite{3}), $\\Omega_{CDM} = 0.94$, $h = 0.5$, no cosmological gravitational waves.}). Currently there are a lot of experimental data (such as the spatial distributions of galaxies, clusters of galaxies and quasars, bulk velosities, CMB anisotropy, and others) which can be used to reconstruct the density perturbation spectrum. Characteristic scales of data are different and vary from $\\sim 10$ Mpc which is a scale of nonlinearity to the horizon scale. However, it seems now the most crucial tests are large-scale CMB anisotropy and the number of galaxy clusters in {\\it top-hat} sphere with radius $R = 8 h^{-1}$ Mpc = 16 Mpc. The former can be easy related to the amplitude of density perturbations through the SW effect (Sachs \\& Wolfe\\cite{4}): $$ \\frac{\\Delta T}{T}(\\vec e) = \\frac{H_0^2}{2(2\\pi)^{3/2}} \\int \\limits_{-\\infty}^{\\infty} \\frac{1}{k^2} \\delta_{\\vec k} e^{i\\vec k\\vec x}d^3\\vec k,\\;\\;\\;\\; \\vec x \\simeq \\frac{2\\vec e}{H_0}, $$ where $\\delta_{\\vec k}$ is a Fourrier transform of density contrast $\\delta(\\vec x) \\equiv \\delta \\rho / \\rho$, $H_0$ is the Hubble constant, $\\langle \\delta_{\\vec k} \\delta_{\\vec k'} \\rangle = P(k) \\delta(\\vec k -\\vec k')$, $P(k) = A k^{n_S} T^2 (\\Omega_\\nu,k)$ is a power spectrum of density perturbations, $A $ is the normalization constant, $T(\\Omega_\\nu,k)$ is a transfer function. The latter determines the value of biasing parameter $b^{-1} \\equiv \\sigma_R$ for spatially flat Universe: $$ \\sigma_R^2 = \\frac{1}{(2\\pi)^3}\\int_{-\\infty}^\\infty P(k) W^2(kR)d^3 \\vec k, $$ where $W(kR) = \\frac{3}{(kR)^3}(\\sin kR - kR\\cos kR)$ is the Fourrier transform of hop-hat window function. Obviously, both normalizations are model-dependent, the $\\Delta T/T$ normalization depends on the amplitude of cosmological gravitational wave spectrum on large scale and, therefore, is related to the model of inflation, the $\\sigma_R$ normalization does depend on the nature of dark matter. Here we prefer to fix $\\sigma_{16}$ to consider the relative contribution of gravitational waves at COBE scale T/S as an additional calculable parameter (instead of considering some inflationary model). Below, we report results based on P\\&S formalism which deals with abundance of gravitationally bounded halos of dark matter. ", "conclusions": "" }, "9803/astro-ph9803058_arXiv.txt": { "abstract": "We apply a unique gas fraction estimator to published X-ray cluster properties and compare the derived gas fractions of observed clusters to simulated ones. The observations are consistent with a universal gas fraction of $0.15\\pm 0.01 h_{50}^{-3/2}$ for the low redshift clusters that meet our selection criteria. The fair sampling hypothesis states that all clusters should have a universal constant gas fraction for all times. Consequently, any apparent evolution would most likely be explained by an incorrect assumption for the angular-diameter distance relation. We show that the high redshift cluster data is consistent with this hypothesis for $\\Omega_0<0.63 $ (95\\% formal confidence, flat $\\Lambda$ model) or $\\Omega_0<0.60$ (95\\% formal confidence, hyperbolic open model). The maximum likelihood occurs at $\\Omega_0=0.2$ for a spatially flat cosmological constant model. ", "introduction": "It has been proposed that clusters of galaxies should be a fair sample of baryonic matter (White \\etal 1993). Rich clusters form through the gravitational collapse of the matter within a 15-30 Mpc diameter volume, where the force of gravity acts equally on all non-relativistic forms of matter. The richest clusters have temperatures above 10 keV, which corresponds to velocity dispersions in excess of 1000 km/sec. Most non-gravitational processes do not appear to affect the bulk of matter at comparable energies, and so we would expect the gas and dark matter to collapse into objects where they are fairly represented. This is confirmed in simulations using a large variety of techniques (Frenk \\etal 1998), where it is found that within the virial radius the gas and dark matter are indeed equally represented with deviations of only 10\\%. Since clusters of galaxies are observed at cosmological distances and are spatially resolved objects, this opens the possibility of directly measuring angular diameter distances if the gas to dark matter ratio were known in advance (Pen 1997). In this paper we compare observed and simulated cluster properties and we estimate the errors in our methods. ", "conclusions": "Comparing a catalog of local cluster properties with simulations, we find that the data is consistent with a universally constant gas fraction of $f_g=0.15\\pm 0.01 h_{50}^{-3/2}$. With the present day uncertainty in $0.025 \\lesssim \\Omega_b h_{50}^2 \\lesssim 0.1$ (Schramm and Turner 1997) and some errors in the Hubble constant, we obtain no useful constaint on $\\Omega_0$ using the low redshift clusters. Independent conservation of baryons and dark matter, however, allows us to constrain $\\Omega_0$ from the apparent evolution of the cluster gas fraction. We have shown that the 3 clusters with measured gas fractions at $z>0.5$ are inconsistent with an $\\Omega=1$ universe and the fair sampling hypothesis. For a spatially flat cosmological constant dominated universe, we obtain a bound of $\\Omega_0<0.63$ (95\\% formal confidence) with a best fit value of $\\Omega_0=0.2$. For a spatially hyperbolic universe with only matter, we find $\\Omega_0<0.60$ (95\\% formal confidence) with the maximum likelihood at $\\Omega_0=0$. The errors are dominated by the temperature uncertainties in the high redshift clusters, and future observations could reduce the errors by a factor of two. This work was supported by the National Science Foundation through REU grant AST 9321943, NASA ATP grant TBD and the Harvard Milton Fund. Computing time was provided by the National Center for Supercomputing Applications. We would like to thank Bill Forman and Christine Jones for providing the cluster X-ray tables." }, "9803/astro-ph9803022_arXiv.txt": { "abstract": "The issue of the approximate isotropy and homogeneity of the observable universe is one of the major topics in modern Cosmology: the common use of the Friedmann--Robertson--Walker [FWR] metric relies on these assumptions. Therefore, results conflicting with the ``canonical'' picture would be of the utmost importance. In a number of recent papers it has been suggested that strong evidence of a fractal distribution with dimension $D\\simeq2$ exists in several samples, including Abell clusters [ACO] and galaxies from the ESO Slice Project redshift survey [ESP]. Here we report the results of an independent analysis of the radial density run, $N( 5in {\\begin{center} \\parbox{5in}{\\footnotesize\\smalllineskip Fig.~\\thefigure. #1} \\end{center}} \\else {\\begin{center} {\\footnotesize Fig.~\\thefigure. #1} \\end{center}} \\fi} \\newcommand{\\tcaption}[1]{ \\refstepcounter{table} \\setbox\\@tempboxa = \\hbox{\\footnotesize Table~\\thetable. #1} \\ifdim \\wd\\@tempboxa > 5in {\\begin{center} \\parbox{5in}{\\footnotesize\\smalllineskip Table~\\thetable. #1} \\end{center}} \\else {\\begin{center} {\\footnotesize Table~\\thetable. #1} \\end{center}} \\fi} \\def\\@citex[#1]#2{\\if@filesw\\immediate\\write\\@auxout {\\string\\citation{#2}}\\fi \\def\\@citea{}\\@cite{\\@for\\@citeb:=#2\\do {\\@citea\\def\\@citea{,}\\@ifundefined {b@\\@citeb}{{\\bf ?}\\@warning {Citation `\\@citeb' on page \\thepage \\space undefined}} {\\csname b@\\@citeb\\endcsname}}}{#1}} \\newif\\if@cghi \\def\\cite{\\@cghitrue\\@ifnextchar [{\\@tempswatrue \\@citex}{\\@tempswafalse\\@citex[]}} \\def\\citelow{\\@cghifalse\\@ifnextchar [{\\@tempswatrue \\@citex}{\\@tempswafalse\\@citex[]}} \\def\\@cite#1#2{{$\\null^{#1}$\\if@tempswa\\typeout {IJCGA warning: optional citation argument ignored: `#2'} \\fi}} \\newcommand{\\citeup}{\\cite} \\def\\@refcitex[#1]#2{\\if@filesw\\immediate\\write\\@auxout {\\string\\citation{#2}}\\fi \\def\\@citea{}\\@refcite{\\@for\\@citeb:=#2\\do {\\@citea\\def\\@citea{, }\\@ifundefined {b@\\@citeb}{{\\bf ?}\\@warning {Citation `\\@citeb' on page \\thepage \\space undefined}} \\hbox{\\csname b@\\@citeb\\endcsname}}}{#1}} \\def\\@refcite#1#2{{#1\\if@tempswa\\typeout {IJCGA warning: optional citation argument ignored: `#2'} \\fi}} \\def\\refcite{\\@ifnextchar[{\\@tempswatrue \\@refcitex}{\\@tempswafalse\\@refcitex[]}} \\def\\pmb#1{\\setbox0=\\hbox{#1} \\kern-.025em\\copy0\\kern-\\wd0 \\kern.05em\\copy0\\kern-\\wd0 \\kern-.025em\\raise.0433em\\box0} \\def\\mbi#1{{\\pmb{\\mbox{\\scriptsize ${#1}$}}}} \\def\\mbr#1{{\\pmb{\\mbox{\\scriptsize{#1}}}}} \\def\\fnm#1{$^{\\mbox{\\scriptsize #1}}$} \\def\\fnt#1#2{\\footnotetext{\\kern-.3em {$^{\\mbox{\\scriptsize #1}}$}{#2}}} \\def\\fpage#1{\\begingroup \\voffset=.3in \\thispagestyle{empty}\\begin{table}[b]\\centerline{\\footnotesize #1} \\end{table}\\endgroup} \\def\\runninghead#1#2{\\pagestyle{myheadings} \\markboth{{\\protect\\footnotesize\\it{\\quad #1}}\\hfill} {\\hfill{\\protect\\footnotesize\\it{#2\\quad}}}} \\headsep=15pt \\font\\tenrm=cmr10 \\font\\tenit=cmti10 \\font\\tenbf=cmbx10 \\font\\bfit=cmbxti10 at 10pt \\font\\ninerm=cmr9 \\font\\nineit=cmti9 \\font\\ninebf=cmbx9 \\font\\eightrm=cmr8 \\font\\eightit=cmti8 \\font\\eightbf=cmbx8 \\font\\sevenrm=cmr7 \\font\\fiverm=cmr5 \\newtheorem{theorem}{\\indent Theorem} \\newtheorem{lemma}{Lemma} \\newtheorem{definition}{Definition} \\newtheorem{corollary}{Corollary} \\newcommand{\\proof}[1]{{\\tenbf Proof.} #1 $\\Box$.} \\textwidth=5truein \\textheight=7.8truein \\def\\qed{\\hbox{${\\vcenter{\\vbox{\t\t\t% \\hrule height 0.4pt\\hbox{\\vrule width 0.4pt height 6pt \\kern5pt\\vrule width 0.4pt}\\hrule height 0.4pt}}}$}} \\renewcommand{\\thefootnote}{\\fnsymbol{footnote}}\t% \\begin{document} \\runninghead{Anisotropy in the Propagation of Radio Wave Polarizations $\\ldots$} { Anisotropy in the Propagation of Radio Wave Polarizations $\\ldots$} \\normalsize\\textlineskip \\thispagestyle{empty} \\setcounter{page}{1} \\copyrightheading{}\t\t\t% \\vspace*{0.88truein} \\fpage{1} \\centerline{\\bf ANISOTROPY IN THE PROPAGATION OF RADIO POLARIZATIONS } \\vspace*{0.035truein} \\centerline{\\bf FROM COSMOLOGICALLY DISTANT GALAXIES} \\vspace*{0.37truein} \\centerline{\\footnotesize PANKAJ JAIN} \\vspace*{0.015truein} \\centerline{\\footnotesize\\it Physics Department, I.I.T. Kanpur} \\baselineskip=10pt \\centerline{\\footnotesize\\it Kanpur, India - 208016} \\vspace*{10pt} \\centerline{\\footnotesize JOHN P. RALSTON} \\vspace*{0.015truein} \\centerline{\\footnotesize\\it Department of Physics and Astronomy, Kansas University} \\baselineskip=10pt \\centerline{\\footnotesize\\it Lawrence, KS-66045, USA} \\vspace*{0.225truein} \\publisher{(received date)}{(revised date)} \\vspace*{0.21truein} \\abstracts{ Radiation traversing the observable universe provides powerful ways to probe anisotropy of electromagnetic propagation. A controversial recent study claimed a signal of dipole character. Here we test a new and independent data set of 361 points under the null proposal of {\\it statistical independence} of linear polarization alignments relative to galaxy axes, versus angular positions. The null hypothesis is tested via maximum likelihood analysis of best fits among numerous independent types of factored distributions. We also examine single-number correlations which are parameter free, invariant under coordinate transformations, and distributed very robustly. The statistics are shown explicitly not to depend on the uneven distribution of sources on the sky. We find that the null proposal is not supported at the level of less than 5\\% to less than 0.1\\% by several independent statistics. The signal of correlation violates parity, that is, symmetry under spatial inversion, and requires a statistic which transforms properly. The data indicate an axis of correlation, on the basis of likelihood determined to be $[{\\rm R.A.}=(0^{\\rm h},9^{\\rm m}) \\pm (1^{\\rm h},0^{\\rm m})$, ${\\rm Decl.} = -1^o\\pm 15^o]$. }{} {} \\vspace*{1pt}\\textlineskip \\textheight=7.8truein \\setcounter{footnote}{0} \\renewcommand{\\thefootnote}{\\alph{footnote}} ", "introduction": "\\noindent The orientation of linear radio polarizations emitted by cosmologically distant galaxies has a consistent relation with the galaxy symmetry axis. Exceedingly small physical effects accumulate during propagation, which conventional measurements can directly probe. Thus electromagnetic radiation traversing the observable universe can detect subtle forms of cosmological anisotropy. A signal with dipole character was claimed recently$^1$ from an analysis of published radio data. Analysis found an ``anisotropy axis\" $\\vec s_{\\rm NR}=(21^{\\rm h}\\pm 2^{\\rm h}, 0^o\\pm 20^o$) governing orientation of polarization of the radio signals varying in a coherent way across the dome of the sky. The origin of this behavior is not clear, and may or may not indicate a fundamental anisotropy on a scale larger than previously found in cosmology. There is a long history of puzzling observations. Beginning in the 1960's observers noticed that Faraday-subtracted polarizations were distributed in peculiar ways relative to the source axes. In 1982 Birch$^2$ empirically observed a coherent angular anisotropy in the off-sets of the polarization and galaxy axes, using a data set of 137 points. Birch's statistical methods were questioned, but more sophisticated studies$^{3,4}$ confirmed surprisingly strong signals in Birch's data. The statistics were not consistent with isotropy at 99.9\\% and 99.98\\% confidence levels, respectively. One of the same groups$^{4,5}$ went on to create an independent set of 277 points and simultaneously introduced a different statistical measure. They obtained no signal in this set and dismissed Birch's results. This left unresolved the puzzling fact that his data had contained a signal at such a high level of statistical significance. When Nodland and Ralston,$^1$ initially unaware of Birch's$^2$ work, independently found a statistically significant signal in an independent set of 160 points, criticisms focused on proposing different statistical baselines$^{6,7,8}$ and again claimed to find no signal of anisotropy. The question of systematic bias in such data had been raised by the authors$^1$ (henceforth $NR$) and earlier$^9$ regarding Birch. Here we report analysis of a considerably larger data set which contains 361 points. We have taken into account criticisms and experience from earlier work, and used the most robust statistical methods available. New progress has been made by paying close attention to the symmetries of the problem. The usual expectation of {\\it independence} of the polarization and sky angular coordinates, or ``uncorrelated isotropy'', happens to represent a definite symmetry, which is that the distribution factors. The classic scientific method becomes applicable: we can test isotropy as a clean hypothesis and see if it can be ruled out, which is immensely powerful. We use generic methods to represent the correlations, emphasising the symmetry that they are {\\it odd} in the polarization variable at hand, which is a consequence of parity (spatial inversion) symmetry.$^{10}$ This simple point resolves many apparent discrepancies between the previous studies. Rather than being at odds with one another, all the facts are now found to be consistent; we know of nothing in contradiction to our conclusions. The data collects variables from cosmologically distant galaxies, as compiled in the literature.$^{2,4,5,11,12,13}$ The data set by $NR$ reproduced that of Carroll et al$^{13}$ except for a half-dozen corrections from the original literature. The compilation of Eichendorf and Reinhardt,$^{11,12}$ available on the NASA-ADC archives, contain numerous sources for which the position angle of the source is listed. We obtained the polarization angle for these sources from Simard-Normandin et al$^{14}$ for all the sources for which they were available. We compiled a total of 152 data points in this fashion. Taking these as our primary data set we added any distinct data points contained in Bietenholz$^5$, making a total of 313 points. Data points were regarded as distinct if they had different Right Ascension, and differed in Declination by more than one degree (which can be attributed to change in convention). This set was further combined with the $NR$ and remaining distinct points of the Birch data, in that order, making a total of 361 data points. In combining these different data sets, we verified that the polarization off-set values for points with coincident Right Ascension and Declination did not differ by more than a few degrees for most of the data. Specifically we found that the disagreement exceeded $5^o$ only for very few points, which if deleted made no difference to our final results. We also verified consistency using a newer 1988 compilation by Broten et al.$^{15}$ The only exception to this rule was found for Birch's data: here the disagreement with other compilations was found to be larger, but still tolerable. All results we report are consistent, and no combination of any large set gave results significantly different from any other. The absence of information available to us on Birch's $RM$ values, plus the possibility of discrepancies in that data, led us to give results both with and without Birch's data. In Figure 1 we show the angular distribution of data, which naturally is not isotropic due to the zone of avoidance and dominance of Northern Hemisphere measurements. We will exhaustively show that the angular distribution is not an issue and cannot be confused with correlation. \\medskip \\begin{figure}[t,b] \\hbox{\\hspace{0em} \\hbox{\\psfig{file=anisa.ps,height=8cm}}} \\caption{ Aithoff-Hammer equal-area plots of the distribution of sources on the dome of the sky, in the equatorial coordinates of the data used. The distribution is somewhat non-uniform due to the zone of avoidance and dominance of Northern Hemisphere measurements. (a) The distribution of the full data set of 332 points, excluding the 29 extra points contained in Birch's compilation. Adding Birch's data makes the set even more uniform. (b) The distribution of the same data set after the cut on rotation measure, $|RM-\\overline{RM}| >6$. Any non-uniformity of the angular distribution is taken into account in all statistics reported. } \\end{figure} \\medskip The observables listed for galaxy $i$ include a major axis orientation angle $\\psi_i$, a linear polarization angle $\\chi_i$, and the angular coordinates of the galaxies on the sky. Other variables may include a resolution parameter, degree of polarization, and the Faraday rotation measure $RM$. The rotation measure is the slope of plots of measured polarization angle versus wavelength-squared. This is known to measure intervening magnetized plasma parameters. A-priori, $RM$ has nothing to do with the variable $\\chi$, which is the polarization angle after Faraday rotation is subtracted. However we have retained this variable, which seems to be informative. Consistent with restricting the study to uncorrelated isotropy, we integrate over the redshift, which happens to be incomplete in the data set in any event. We let $\\beta=\\chi-\\psi$ be the angle between the plane of polarization and the symmetry axis of the source. The variables $ \\chi$ and $\\psi$ are determined up to a multiple of $\\pi$; $\\beta$ runs from $-\\pi$ to $\\pi$.$^{10}$ To deal with the $\\pi$ ambiguity of polarization and axis measurements, one can map $\\beta \\rightarrow Y(\\Omega)$, where $\\Omega$ is a variable defined on twice the interval. A popular map is ``{\\it Map 1}'', $\\Omega_1(\\beta) = 2 \\beta$. The function $Y$ is represented by a Fourier series with periodicity $2\\pi$, assuring that the transformation $\\beta \\rightarrow \\beta^\\prime = \\beta \\pm \\pi$ leaves $Y(\\Omega)$ invariant. The first Fourier components create a 2-component vector-like object $\\vec Y(\\Omega) = (\\cos(\\Omega), \\sin(\\Omega))$. When the components of $\\vec Y(\\Omega)$ are used in statistical analysis, there is naturally a Jacobian factor which represents the choice of {\\it Map}. By no means, then, is {\\it Map 1} sacred, and other maps are discussed below. The angular positions on the dome of the sky are mapped into their 3-dimensional Cartesian vector positions $\\vec X$ on a unit sphere. Since we do not model this distribution, but take it from the data, this standard map is adequate. When coordinate origins are changed, the components of $\\vec X$ transform by standard rules; one can go on to make nicely transforming distributions and tensor correlations. The two choices measuring $\\chi$ relative to $\\psi$ or $\\psi+\\pi/2$ correspond to $\\vec Y\\rightarrow -\\vec Y$. This does not mix the 2 components of $\\vec Y$, which will be called ``even'' (for $\\cos(\\Omega)$) and ``odd'' (for $\\sin(\\Omega)$) following the transformation property of being even or odd, respectively, under parity (spatial inversion). As discussed in detail elsewhere,$^{10}$ functions of the offset angles have the corresponding parity if they are even or odd functions of $\\beta$, as intuitively evident from the handed ``sense of twist'' a parity-odd quantity conveys. The invariant correlations discussed below avoid any question of coordinate origin (either in polarization quantities or in angular positions on the dome of the sky) by being totally independent of the choice of angular origin. The standard assumption of statistical independence corresponds to a distribution $g(\\Omega,\\vec X)=h(\\Omega) f(\\vec X )$. This is a very broad class of distributions, with $h(\\Omega)$ and $f(\\vec X )$ completely unrestricted, which nevertheless has symmetries allowing it to be tested. All statistics will be compared to baselines using the actual distribution of the data $f(\\vec X)$ on the dome of the sky in Monte Carlo simulations. Statistics based on assuming independence of polarizations and positions will be compared with a simple correlated ansatz of the form $h(\\Omega) C(\\Omega, \\vec X) f(\\vec X)$. The case $C=1$ reduces to the uncorrelated case. ", "conclusions": "\\noindent In presenting a study of restricted scope, our conclusions are most crisply phrased in a negative sense: {\\it the null hypothesis of uncorrelated isotropy is not supported. On the basis of significance, it can be ruled out}. By the nature of this study, one is constrained from concluding prematurely what the correlation found may represent. Under many separate statistical probes, the evidence against isotropy in the data is significant at $95\\%-99.99\\%$ (roughly $2-4\\sigma$) confidence levels. This is not the first such finding, but just one more among a number of studies accumulating over the years. While no evidence of systematic bias is found, we strongly reiterate the possibility. Yet the persistence of the effect seems to indicate physical processes outside the framework which has been used to interpret the data conventionally. Associated with this behavior are persistent axis parameters concordant with the axis parameters found in Nodland and Ralston,$^1$ and which subsequently have been found to coincide with the CMB dipole direction.$^{20,21}$ Nevertheless this is a new field and it would be premature to fix on a physical origin now. We therefore postpone more detailed conclusions, and recommend that physical models be used to suggest suitable directions of research. Local effects, while traditionally held to be under control, can potentially be ruled out with redshift information. Resources exist to generate cosmological radio data sets with many more points, and the time may be ripe for clever technological advances that could be revolutionary. New analysis combined with new data might tell us what is causing the effect. \\bigskip \\nonumsection{Acknowledgements} \\noindent We thank Borge Nodland, Hume Feldman, Doug McKay and G. K. Shukla for useful comments. Supported by DOE grant number 85 ER40214, the KU General Research Fund, the NSF-K*STAR Program under the Kansas Institute for Theoretical and Computational Science and DAE grant number DAE/PHY/96152. \\nonumsection{References} \\noindent" }, "9803/astro-ph9803346_arXiv.txt": { "abstract": "We present high-resolution imaging of the young binary, T Tauri, in continuum emission at $\\lambda$=3~mm. Compact dust emission with integrated flux density 50 $\\pm$ 6 mJy is resolved in an aperture synthesis map at 0\\farcs5 resolution and is centered at the position of the optically visible component, T Tau N\\null. No emission above a 3$\\sigma$ level of 9 mJy is detected 0\\farcs7 south of T Tau N at the position of the infrared companion, T~Tau~S\\null. We interpret the continuum detection as arising from a circumstellar disk around T~Tau~N and estimate its properties by fitting a flat-disk model to visibilities at $\\lambda$ = 1 and 3~mm and to the flux density at $\\lambda$ = 7~mm. Given the data, probability distributions are calculated for values of the free parameters, including the temperature, density, dust opacity, and the disk outer radius. The radial variation in temperature and density is not narrowly constrained by the data. The most likely value of the frequency dependence of the dust opacity, $\\beta$ = $0.53^{+0.27}_{-0.17}$, is consistent with that of disks around other single T Tauri stars in which grain growth is believed to have taken place. The outer radius, R = 41$^{+26}_{-14}$~AU, is smaller than the projected separation between T Tau N and S, and may indicate tidal or resonance truncation of the disk by T~Tau S\\null. The total mass estimated for the disk, log(M$_D$/M$_\\odot$) = ${-2.4}^{+0.7}_{-0.6}$, is similar to masses observed around many single pre--main-sequence sources and, within the uncertainties, is similar to the minimum nebular mass required to form a planetary system like our own. This observation strongly suggests that the presence of a binary companion does not rule out the possibility of formation of a sizeable planetary system. {\\it Subject headings:} circumstellar matter --- stars:pre-main sequence --- star:individual (T Tauri) ", "introduction": "Observations of T~Tau at $\\lambda$ = 2.8 mm were carried out from 1996 December to 1997 March with the 9-element BIMA array. Continuum emission was measured in an 800 MHz bandwidth centered at 108 GHz. Data were taken with the array in the A, B, and C configurations, providing baseline coverage from 2.1 k$\\lambda$ to 420 k$\\lambda$ with sensitivity to emission on size scales up to $\\sim$60\\arcsec. Interleaved observations of a nearby quasar, 0431+206, were included as a check on the phase de-correlation on long baselines. The phase calibrator was 0530+135, with an assumed flux of 3.1~Jy during A array. Data were calibrated and mapped using the Miriad package. A map of 0431+206 yielded an image of a point source, indicating little atmospheric degradation of the resolution in the observations of T~Tau. An aperture synthesis image of T Tau was constructed with data from the A array alone (the longest baselines) and is displayed in Figure \\ref{map}. The beam size is $0\\farcs59 \\times 0\\farcs39$ at a position angle of 48\\arcdeg. It is clearly evident that the compact 3~mm emission arises from circumstellar material surrounding T~Tau~N only. Peak emission of 32 mJy beam$^{-1}$ is centered at RA(J2000) 04:21:59.424 and Dec(J2000) 19:32:06.41 with an absolute positional uncertainty of $\\sigma$ = $\\pm$0\\farcs07 as determined from observations of 0431+206. The peak position is within 1.6$\\sigma$ of the Hipparcos coordinates for T~Tau~N, but 7.8$\\sigma$ distant from the position of T~Tau~S\\null. The emission is resolved with an approximate deconvolved source size of 0\\farcs45 $\\times$ 0\\farcs32 (FWHM) (63 $\\times$ 45~AU) at PA 19\\arcdeg, assuming an elliptical Gaussian shape. The integrated flux density, 50 $\\pm$ 6 mJy (where the uncertainty is dominated by the flux calibration error), is consistent with the 100 GHz value measured at lower resolution by Momose et al.\\ (1996), 48 $\\pm$ 7 mJy. No emission is detected at the position of T~Tau~S above the 3$\\sigma$ upper limit of 9 mJy beam$^{-1}$, and no circumbinary emission is apparent. The integrated flux density detected in the compact C array is no greater than that for A array. Since the C array observations are most sensitive to emission on larger spatial scales, up to 60$''$, the absence of detectable excess emission implies that the majority of continuum emission detected in a 1$'$ aperture at $\\lambda$ = 3 mm originates from a circumstellar region around T Tau N\\null. The 3$\\sigma$ upper limit on circumbinary emission, accounting for thermal noise and flux calibration errors, is 17 mJy. Our measurement of the 108 GHz flux density from T~Tau~N is plotted in Figure \\ref{sed} together with high-resolution measurements at 43 GHz ($\\lambda$ = 7~mm) (Koerner et al.\\ 1998), 267 GHz ($\\lambda$ = 1~mm) and 357 GHz ($\\lambda$ = 0.8~mm) (Hogerheidje et al.\\ 1997). All four points can be fitted by a single power-law curve with $\\chi^2$ = 1.8 and goodness of fit 0.4. The best-fit spectral index $\\alpha$=d(log($F_\\nu$))/d(log($\\nu$)) is 2.30$\\pm$0.1. This value is in good agreement with those from disks around single T Tauri stars and consistent with thermal emission from large circumstellar dust grains (Beckwith \\& Sargent 1991; Mannings \\& Emerson 1994; Koerner et al.\\ 1995), suggesting that the continuum emission from T Tau N has a single origin in thermal dust emission along the entire wavelength range from $\\lambda$ = 1 to 7~mm. We argue that the 3mm emission arises from a circumstellar disk around T Tau N\\null. The spectral slope is consistent with radiation from dust grains (see Koerner et al.\\ 1998 for a more stringent limit on the possible contribution of free-free or gyrosynchrotron emission). The star is detected optically, even though the lower limit to the dust mass is high; if the dust were distributed in a uniform density sphere with a size given by the A array fit, the extinction to the star would be A$_V \\sim$ 1000. \\section {Circumstellar disk models for T~Tau~N} \\subsection{Model description} To refine estimates of the parameters of dust around T~Tau~N, we fit the visibility amplitudes directly with a model of a circumstellar disk. While computationally intensive, fitting in the visibility plane rather than the image plane avoids the non-linear process of deconvolution and allows the instrumental errors to be included in a consistent manner. Our disk model is thin and circularly symmetric with a flux from an annular region at radius $r$ given by \\begin{equation} dS_{\\nu}(r) = {2 \\pi \\cos\\theta \\over D^2 } B_{\\nu} (1 - e^{-\\tau}) r \\, dr, \\label{disk_flux} \\end{equation} where $B_{\\nu}$ is the Planck function, $D$ is the distance, and $\\theta$ the inclination angle. Flared disks have been invoked to explain the flat infrared spectral indices of some T Tauri stars (Kenyon \\& Hartmann 1987); however, the effects due to flaring are not important at these long wavelengths (Chiang \\& Goldreich 1997). The visibility amplitude $V$ at $uv$-distance $\\eta$ is calculated with a Hankel transform, \\begin{equation} V(\\eta) = 2\\pi \\int S_{\\nu}(r)J_0(2\\pi\\eta r/D)r\\,dr, \\label{hankel} \\end{equation} where $J_0$ is the Bessel function. The inner radius is set by the dust destruction temperature (2000~K); the exact value has little effect on the millimeter flux density. The temperature, surface density and dust opacity are described by power-law relations: $ T(r) = T_{\\rm 10~AU} (r/{\\rm 10~AU})^{-q},\\ \\Sigma (r) = \\Sigma_{\\rm 10~AU} (r/{\\rm 10~AU})^{-p},$ and $\\kappa_{\\nu} = \\kappa_o (\\nu/\\nu_o)^{\\beta}$. Due to its dependence on the product of surface density and dust opacity, the optical depth scales with the corresponding power-law exponents in radius and frequency, \\begin{equation} \\tau (r,\\nu) = {\\Sigma \\kappa_{\\nu} \\over \\cos\\theta} \\equiv \\tau_{\\rm 10~AU} \\Big({r \\over {\\rm 10~AU}}\\Big)^{-p}\\Big({\\nu \\over \\nu_o}\\Big)^{\\beta}. \\end{equation} The reference value for $\\kappa_o$ is 0.1~g$^{-1}$~cm$^{2}$ at $\\nu_o = 1200$ GHz ($\\lambda$ = 250~$\\mu$m; Hildebrand 1983). Eight parameters were varied in the model-data comparison: $T_{\\rm 10~AU}$, $ q$, $\\tau_{\\rm 10~AU},$ $p,$ $\\beta,$ $r_{out},$ $\\theta$, $\\mbox{and}\\ \\alpha$, the position angle of the disk on the sky. In addition to the 3~mm visibilities, those measured at 1~mm by Hogerheidje et al.\\ (1997) were used together with the 7~mm flux density (Koerner et al.\\ 1998). The IRAS 100 $\\mu$m flux of 120 Jy (Strom et al.\\ 1989) was included as an upper limit. For each value of $\\theta$ and $\\alpha$, the $u$ and $v$ coordinates for each visibility were de-projected to those of a face-on disk, then binned in annuli of de-projected $uv$-distance for comparison to model values calculated by Eqn.\\ \\ref{hankel}. Over 5 million models were compared to the data within an 8-dimensional grid in parameter space. Logarithmic grid spacings were used for the temperature, optical depth and outer radius. To quantitatively assess the reliability of estimates of properties of a disk around T Tau N, we calculate the probability distribution for each parameter over the entire range of models. A detailed description of this Bayesian approach is given in Lay et al.\\ (1997). Given the data, the probability of a model with a particular set of parameter values is proportional to $e^{-\\chi^2}$ where $\\chi^2$ is the standard squared difference between data and model, weighted by the uncertainty in the data. A systematic error in overall flux calibration was accounted for by normal weighting of a range of model flux scalings with 1$\\sigma$ corresponding to a 10\\% difference in flux. The final probability for a given model was taken as the sum of probabilities over all flux scalings and multiplied by a factor of $\\sin \\theta$ to account for the fact that edge-on disks are more likely than face-on in a randomly oriented sample. Finally, the relative likelihood of each parameter value was calculated by adding the probabilities of all models with that parameter value. \\subsection{Parameter results} The resulting probability distributions for the disk parameters are given below. The ranges considered for $T_{\\rm 10~AU}$, $\\tau_{\\rm 10~AU}$, and $r_{out}$ were sufficiently wide to bracket all values with significant probability. Parameters values quoted are at the median of the probability distribution with an error range encompassing 68\\% of the total probability. However, it is important to keep in mind that the distribution is not Gaussian. The disk outer radius is $r_{out} = 41^{+26}_{-14}$~AU (Figure \\ref{prob}). The probability that it exceeds the projected binary separation (100 AU) is only 3\\%. Models with large outer radii generally have higher values of $p$ and lower values of $\\tau_{\\rm 10~AU}$. The value for the dust mass opacity index, $\\beta = 0.53^{+0.27}_{-0.17}$ (Figure \\ref{prob}), is consistent with values measured for circumstellar disks around T Tauri stars (Beckwith \\& Sargent 1991; Koerner et al.\\ 1995). There is an 85\\% probability that $\\beta$ $\\ge 0.30$, the value calculated by assuming optically thin emission (S$_\\nu \\propto \\nu^{2 + \\beta}$) and fitting a straight line to the 7, 3, and 1~mm fluxes. As discussed below, this implies that at least some of the emission is optically thick. There is some correlation between models with high values of $\\beta$ and those with high $\\tau$ and low $T$. The temperature, $T_{\\rm 10~AU} = 26^{+34}_{-13}$~K, and $\\lambda$ = 3~mm optical depth, $\\tau_{\\rm 10~AU} = 0.50^{+0.83}_{-0.36}$, are not as tightly constrained as $r_{out}$. Although many models have $\\tau > 1$ at 10~AU, most (89\\%) radiate more than half their total emission in an optically thin regime. Note that for an optically thin disk in the Rayleigh-Jeans regime, Eqn.\\ \\ref{disk_flux} becomes degenerate in $T$ and $\\tau$. Consequently, temperature and optical depth are anti-correlated for $\\tau <$ 0.5 at 10~AU. The data do not narrowly constrain values for $q$, $p$, $\\theta$ or $\\alpha$. The ranges used for $p$ and $q$ were $p$=0.5--2.0 and $q$=0.4--0.75. Steeper density profiles are slightly favored over shallow ones; the probability that $p$ is $\\ge$ 1.5 is 65\\%. The lack of preferred values for $p$ and $q$ is largely due to the degeneracy of $p$ and $q$ for optically thin emission. It may be possible to constrain the disk parameters further by including data from additional wavelengths. Mid-infrared flux densities are often used to determine $q$ and $T_o$, for example. We chose not to include these data, however, because the mid-infrared flux traces material within a few AU of the star, while the millimeter data is sensitive mainly to material tens of AU away. The simple power-law relations assumed in the model may not be valid over such a large range of disk radii and physical conditions. The disk model masses, weighted by the model probabilities, were binned to derive a median value log(M$_D$/M$_\\odot$) = ${-2.4}^{+0.7}_{-0.6}$ (Figure \\ref{prob}). The wide range in mass is due largely to the range of $\\tau$ values that fit the data. For a given 3~mm flux, disks with higher mass correspond to those with higher $\\tau$ and $\\beta$. The median value for the disk mass, M$_D$ = $4 \\times 10^{-3}$ M$_\\odot$, is toward the low end of disk masses typically derived for classical T~Tauri stars (e.g.\\ Beckwith et al.\\ 1990). We note, however, a large dependence of mass estimates on values assumed for the other parameters. Beckwith et al.\\ assumed $\\beta$ = 1 and an outer radius of 100~AU\\null. If we consider only models with $\\beta \\ge 0.75$ and an outer radius $>$ 50~AU, the median mass increases to $10^{-2}$ M$_\\odot$, similar to the minimum mass solar nebula. ", "conclusions": "It has been conjectured that the formation of planetary systems arises from the collapse of protostellar clouds which rotate more slowly than clouds from which binaries form (Safronov \\& Ruzmaikina 1985). This, in turn, raises the possibility that planetary systems may fail to form in the binary environment. In contrast, our observations and modeling demonstrate that a substantial mass of material, with size like that of the solar system, can exist around a star in a binary system with separation not less than 100 AU\\null. This material constitutes a large reservoir available to planet-forming processes; the low value of the mass opacity index, $\\beta$ =0.53, further suggests that the formation of larger grains may already be underway as a first step toward planetesimal formation (Beckwith \\& Sargent 1991; Mannings \\& Emerson 1994; Koerner et al.\\ 1995). Our observations rule out a greater circumstellar mass of dust around T Tau S---whether in an edge-on circumstellar disk or compact spherical envelope---as a simple explanation for the origin of increased extinction along the line of sight to T Tau's infrared companion. Due largely to the low optical depth of dust continuum emission at $\\lambda$ = 3mm, however, the true source of extra extinction is not identified unambiguously. Our estimate of the outer radius is smaller than the projected separation between T~Tau~N and S and suggests that a disk around T Tau~N does not obscure T~Tau~S\\null. However, we did not consider models with an exponential density profile at the outer edge like that of Hogerheidje et al.\\ (1997). We note that since $\\kappa(500\\ {\\rm nm})/\\kappa(3\\ \\rm{mm}) \\sim 10^3$--$10^4$ (e.g., Pollack et al.\\ 1994), the material that provides the extinction toward T Tau S could easily be undetectable at 3 mm. Thus, our data are consistent either with obscuration of T Tau S by tenuous outer regions of the T Tau N disk or with truncation of the T Tau N disk by T Tau S as discussed below. If the binary components are in a bound orbit, as suggested by their common proper motion, the distribution of circumstellar material at the outer edge of the disk will be affected by the gravitational influence of the companion. In models of disk/companion interactions, the size of the circumstellar disk depends on the mass ratio of the stars and their separation (Papaloizou \\& Pringle 1977, Artymowicz \\& Lubow 1994). The radius of the circumprimary disk typically ranges from 0.3 to 0.4 times the separation for mass ratios of 1 to 0.3 and circular orbits. If the orbital plane of the system is viewed nearly face-on and the binary separation is 100--110~AU, the tidal disk radius would be 30--40~AU\\null. The disk size estimated from modeling our observations, 27 $< R <$ 67~AU, is consistent with the range of values predicted for tidal truncation. If tidal truncation has occurred and we are seeing the true size of the disk in our maps, then the disk around T Tau N is not obscuring T Tau S\\null. However, high-resolution observations at sub-millimeter wavelengths or in molecular transitions that better trace low-column-density material are needed to adequately solve this problem. Finally, we point out that the disk around T Tau N is similar to those observed around some single low-mass stars, regardless of the reliability of model assumptions that lead to an estimate of the absolute value of its mass and size, since the millimeter-wave flux from T Tauri is among the brightest measured from a large sample of young low-mass stars (cf.\\ Beckwith et al.\\ 1990, Osterloh \\& Beckwith 1995), and the nominal FWHM size of the emission is similar to that of single-star disks for which sufficiently high-resolution observations have been carried out (e.g. Lay et al.\\ 1994; Mundy et al.\\ 1996). If the largest disks around T Tauri stars typically yield planetary systems like our own, it is plausible that the disk around T Tau N will too, in spite of the presence of a companion at a distance of 100 AU or greater. Since the majority of pre-main-sequence stars are in multiple systems, the possibility of planetary formation in binaries like T Tauri invites consideration of the likelihood that a non-negligible fraction of binary stars contribute substantially to the estimated fraction of stars that possess planetary systems." }, "9803/astro-ph9803170_arXiv.txt": { "abstract": "We use recent data obtained by three (OSSE, BATSE, and COMPTEL) of four instruments on board the Compton Gamma Ray Observatory, to construct a model of Cyg X-1 which describes its emission in a broad energy range from soft X-rays to MeV $\\gamma$-rays self-consistently. The $\\gamma$-ray emission is interpreted to be the result of Comptonization, bremsstrahlung, and positron annihilation in a hot optically thin and spatially extended region surrounding the whole accretion disk. For the X-ray emission a standard corona-disk model is applied. We show that the Cyg X-1 spectrum accumulated by the CGRO instruments during a $\\sim$4 year time period between 1991 and 1995, as well as the HEAO-3 $\\gamma_1$ and $\\gamma_2$ spectra can be well represented by our model. The derived parameters match the observational results obtained from X-ray measurements. ", "introduction": "One of the brightest sources in the low-energy $\\gamma$-ray sky, Cyg X-1, has been extensively studied during the last three decades since its discovery (\\cite{Bowyer65}, for a review see \\cite{Oda77,LiangNolan84}). It is a high-mass binary system (HDE~226868) with an orbital period of 5.6~days consisting of a blue supergiant and presumably a black hole (BH) with a mass in excess of $5M_\\odot$ (\\cite{Dolan92}). The separation of the two components is $\\approx4\\times10^{12}$ cm (\\cite{Beall84}). A periodicity of 294~d found in X-ray and optical light curves is thought to be related to precession of the accretion disk (\\cite{Priedhorsky83,Kemp83}). The X-ray flux of Cyg X-1 varies on all observed timescales down to a few milliseconds (e.g., \\cite{Cui97}), but the average flux exhibits roughly a two-modal behaviour. Most of its time Cyg X-1 spends in a so-called `low' state where the soft X-ray luminosity (2--10 keV) is low. The low-state spectrum is hard and can be described by a power-law with a photon index of $\\sim1.7$ in the 10--150 keV energy band. There are occasional periods of `high' state emission, in which the spectrum consists of a relatively stable soft blackbody component and a weak and variable hard power-law component. Remarkable is the anticorrelation between the soft and hard X-ray components (\\cite{LiangNolan84}), which is clearly seen during the transition phases between the two states. Cyg X-1 is believed to be powered by accretion through an accretion disk. Its X-ray spectrum indicates the existence of a hot X-ray emitting and a cold reflecting gas. The soft blackbody component is thought to consist of thermal emission from an optically thick and cool accretion disk (\\cite{ShakuraSunyaev73,Pringle81,Balucinska95}). The hard X-ray part ($\\ga10$ keV) with a break at $\\sim150$ keV has been attributed to thermal emission of the accreting matter Comptonized by a hot corona with temperature from tens to hundred keV (\\cite{SunyaevTitarchuk80,LiangNolan84}). A broad hump peaking at $\\sim20$ keV (\\cite{Done92}), an iron K$\\alpha$ emission line at $\\sim6.2$ keV with an equivalent width $\\sim100$ eV (\\cite{Barr85,Kitamoto90}, see also \\cite{Ebisawa96} and references therein), and a strong iron K-edge (e.g., see \\cite{Inoue89,Tanaka91,Ebisawa92},1996) have been interpreted as signatures of Compton reflection of hard X-rays off cold accreting material. In addition, there have also been sporadic reports of a hard spectral component extending into the MeV region. The most famous one was the so-called `MeV bump' observed at a $5\\sigma$ level during the HEAO-3 mission (\\cite{Ling87}). For a discussion of the pre-CGRO data and $\\gamma$-ray emission mechanisms see, e.g., a review by Owens \\& McConnell (1992). The COMPTEL spectrum accumulated over 15 weeks of real observation time during the 1991--95 time period shows significant emission out to several MeV (\\cite{McConnell97}), which, however, remained always by more than an order of magnitude below the MeV bump reported from the HEAO-3 mission. The annihilation line search provided only tentative (1.9$\\sigma$) evidence for a weak 511 keV line with a flux of $(4.4\\pm2.4)\\times10^{-4}$ photons cm$^{-2}$ s$^{-1}$ (\\cite{LingWheaton89}). Recent OSSE observations (\\cite{Phlips96}) resulted only in upper limits with values of $\\le7\\times10^{-5}$ cm$^{-2}$ s$^{-1}$ for a narrow 511 keV line and $\\le2\\times10^{-4}$ cm$^{-2}$ s$^{-1}$ for a broad feature at 511 keV. Although an unified view for the X-ray spectra of BH candidates and their spectral states has yet to be constructed, the qualitative picture seems to be quite clear. Current popular models include an optically thick disk component, a hot Comptonizing region (e.g., \\cite{Haardt93,Gierlinski97}), and/or an advection-dominated accretion flow (e.g., \\cite{Abramowicz95,NarayanYi95} and references therein). The spectral changes are probably governed by the mass accretion rate (e.g., \\cite{Chen95,Esin97}). \\begin{deluxetable}{lc}% \\tablecolumns{2} \\footnotesize \\setlength{\\tabcolsep}{0.25em} \\tablecaption{ Luminosity of Cyg X-1. \\label{table1}} \\tablewidth{7cm} \\tablehead{ \\colhead{Energy band} & \\colhead{Luminosity, $10^{36}$erg/s} } \\startdata $\\geq0.02$ MeV & $26$ \\nl 0.02--0.2 MeV & $20.5$ \\nl 0.2--1 MeV & $4.8$\\nl $\\geq1$ MeV & $0.6$\\nl \\enddata \\end{deluxetable}% This picture, however, provides no explanation for the observed $\\gamma$-ray emission (e.g., McConnell et al. 1997). The hard MeV tail can not be explained by standard Compton models because they predict fluxes which are too small at MeV energies, and thus another mechanism is required. The models developed so far connect the $\\gamma$-ray emission with a compact hot core ($\\sim400$ keV or more) in the innermost part of the accretion disk, which emits via bremsstrahlung, Compton scattering, and annihilation (\\cite{LiangDermer88,SkiboDermer95}), or with $\\pi^0$ production due to collisions of ions with nearly virial temperature (e.g., \\cite{KolykhalovSunyaev79,JourdainRoques94}). Li, Kusunose \\& Liang (1996) have shown that stochastic particle acceleration via wave-particle resonant interactions in plasmas ($\\sim100$ keV) around the BH could provide a suprathermal electron population, and is able to reproduce the hard state MeV tail. The possibility of Comptonization in the relativistic gas inflow near the BH horizon has been discussed by Titarchuk \\& Zannias (1998). We use the recent data obtained by three of four instruments aboard CGRO to construct a model of Cyg X-1, which describes its emission in a wide energy range from soft X-rays to MeV $\\gamma$-rays (\\cite{Moskalenko97}). Instead of a compact (pair-dominated) $\\gamma$-ray emitting region, we consider an optically thin and spatially extended one surrounding the whole accretion disk. It produces $\\gamma$-rays via Comptonization, bremsstrahlung and positron annihilation. For the X-ray emission the corona-disk model is retained. In section 2 we discuss the combined OSSE--BATSE--COMPTEL spectrum of Cyg X-1. Our model and the inferred results are described in sections 3--4, and the implications are discussed in section 5. The applied formalism is given in the Appendix. ", "conclusions": "The data obtained recently by the CGRO instruments allow us to construct a model of Cyg X-1 which describes its emission from soft X-rays to MeV $\\gamma$-rays self-consistently. This model is based on the suggestion that the $\\gamma$-ray emitting region is a hot optically thin and spatially extended proton-dominated cloud, the outer corona. The emission mechanisms are bremsstrahlung, Comptonization, and positron annihilation. For X-rays a standard corona-disk model is applied. The CGRO spectrum of Cyg X-1 accumulated over a $\\sim$4 years period between 1991 and 1995, as well as the HEAO-3 $\\gamma_1$, and $\\gamma_2$ spectra can be well represented by our model. The derived parameters match also the basic results of the X-ray observations. A fine tuning of the model would require further Monte Carlo simulations and more accurate spectral measurements. In this respect, the solution of the discrepancy between the OSSE and BATSE normalizations would be of particular importance." }, "9803/astro-ph9803200_arXiv.txt": { "abstract": "s{Recent observations of microlensing events in the Large Magellanic Cloud suggest that a sizable fraction of the galactic halo is in the form of Massive Astrophysical Compact Halo Objects (MACHOs). Although the average MACHO mass is presently poorly known, the value $\\sim 0.1 M_{\\odot}$ looks as a realistic estimate, thereby implying that brown dwarfs are a viable and natural candidate for MACHOs. We describe a scenario in which dark clusters of MACHOs and cold molecular clouds (mainly of $H_2$) naturally form in the halo at galactocentric distances larger than 10-20 kpc. Moreover, we discuss various experimental tests of this picture.} \\normalsize\\baselineskip=15pt ", "introduction": "Since 1993 several microlensing events have been detected towards the Large Magellanic Cloud by the MACHO and EROS collaborations. Everybody agrees that this means that Massive Astrophysical Compact Halo Objects (MACHOs) have been discovered. Yet, the specific nature of MACHOs is unknown, mainly because their average mass turns out to depend strongly on the assumed galactic model. For instance, the spherical isothermal model would give $\\sim 0.5 M_{\\odot}$ whereas the maximal disk model would yield $\\sim 0.1 M_{\\odot}$ for that quantity. What can be reliably concluded today is only that MACHOs should lie in the mass range $0.05 M_{\\odot} - 1~M_{\\odot}$. Remarkably enough, the MACHO team has claimed that the fraction of galactic matter in the form of MACHOs is fairly model independent and -- within the present statistics -- should be $\\sim 50 \\%$. What is the most realistic galactic model? Regretfully, no clear-cut answer is presently available. Nevertheless, the current wishdom -- that the Galaxy ought to be best described by the spherical isothermal model -- seems less convincing than before and nowadays various arguments strongly favour a nonstandard galactic halo. Indeed, besides the observational evidence that spiral galaxies generally have flattened halos, recent determinations of both the disk scale length, and the magnitude and slope of the rotation at the solar position indicate that our galaxy is best described by the maximal disk model. This conclusion is further strengthened by the microlensing results towards the galactic centre, which imply that the bulge is more massive than previously thought. Correspondingly, the halo plays a less dominant r\\^ole than within the spherical isothermal model, thereby reducing the halo microlensing rate as well as the average MACHO mass. A similar result occurs within the King-Michie halo models, which also take into account the finite escape velocity and the anisotropies in velocity space (typically arising during the phase of halo formation). Moreover, practically the same conclusions also hold for flattened galactic models with a substantial degree of halo rotation. So, the expected average MACHO mass should be smaller than within the spherical isothermal model and the value $\\sim 0.1~M_{\\odot}$ looks as a realistic estimate to date. This fact is of paramount importance, since it implies that brown dwarfs are a viable and natural candidate for MACHOs. Still -- even if MACHOs are indeed brown dwarfs -- the problem remains to explain their formation, as well as the nature of the remaining dark matter in galactic halos. We have proposed a scenario in which dark clusters of MACHOs and cold molecular clouds -- mainly of $H_2$ -- naturally form in the halo at galactocentric distances larger than $10-20$ kpc (somewhat similar ideas have also been put forward by Ashman and by Gerhard and Silk). Below, we shall review the main features of this model, along with its observational implications. ", "conclusions": "" }, "9803/astro-ph9803085_arXiv.txt": { "abstract": "We compute the polarization of the Ly$\\alpha$ line photons emerging from an anisotropically expanding and optically thick medium, which is expected to operate in many Ly$\\alpha$ emitting objects including the primeval galaxy DLA~2233+131 and Lyman break galaxies. In the case of a highly optically thick medium, the escape of resonance line photons is achieved by a large number of resonant local scatterings followed by a small number of scatterings in the damping wing. We show that some polarization can develop because the wing scatterings are coupled with strong spatial diffusion which depends on the scattering geometry and kinematics. The case of a slab with a finite scattering optical depth and expansion velocity of $\\sim 100~\\kms$ is investigated and it is found that Ly$\\alpha$ photons are emergent with the linear degree of polarization up to 10 per cent when the typical scattering optical depth $\\tau {\\gtrsim} 10^5$. We subsequently investigate the polarization of Ly$\\alpha$ photons emerging from a spherical shell obscured partially by an opaque component and we obtain $\\sim$ 5 per cent of polarization. It is proposed that a positive detection of polarized Ly$\\alpha$ with P-Cygni type profile from cosmological objects can be a strong test of the expanding shell structure obscured by a disk-like component. ", "introduction": "Various astronomical objects in the cosmological scales show P-Cygni type profiles in the Ly$\\alpha$ emission. These objects include the most remote galaxy at $z=4.92$ gravitationally lensed by CL1358+62 \\markcite{fra97} (Franx et al. 1997), high $z$ galaxies observed with the {\\it Hubble Space Telescope} and the Keck telescopes \\markcite{ste96a, ste96b, gia96, low97} (Steidel et al. 1996, Giavalisco et al. 1996, Lowenthal et al. 1997) and the damped Lyman $\\alpha$ (hereafter DLA) candidates \\markcite{djor96, djor97} (Djorgovski et al. 1996, 1997). Similar P-Cygni Ly$\\alpha$ profiles are found in nearby starburst galaxies, which are sometimes classified as Wolf-Rayet galaxies, blue compact galaxies, or H~II galaxies \\markcite{kun96, hec97, sah97, leq95, leg97} (e.g. Kunth et al. 1996, Heckman and Leitherer 1997, Sahu and Blades 1997, Lequeux et al. 1995, Legrand et al. 1997, etc.). The column density $N_{HI}$ of neutral hydrogen in these systems is usually found to be in the range $N_{HI}\\sim 10^{19-21}~\\cm^{-2}$. The primeval galaxies or the first star clusters expected to be found at $z>5$ epoch may possess a central super star cluster surrounded by neutral hydrogen of high column density \\markcite{hai97a, hai97b} (Haiman and Loeb 1997a,b). These surrounding layers can be accelerated by the expanding H~II region just outside the super star cluster. It is hoped that in the near future with the advent of the {{\\it Next Generation Space Telescope} (NGST), the infrared spectra of these infant galaxies will be accessible and that the observational confirmation of the ubiquity of P-Cygni type Ly$\\alpha$ profiles may test the above hypothesis. \\markcite{AL98} Ahn and Lee (1998, hereafter AL98) investigated the Ly$\\alpha$ line formation in a thick and expanding medium. It was emphasized that the profile formation should be studied by accurately computing the contributions from photons back scattered by receding medium and wing-scattered photons \\markcite{leg97} (see also Legrand et al. 1997). It is well known that the properties of the Ly$\\alpha$ photons scattered in the damping wing are characterized by the Rayleigh phase function \\markcite{ste80} (e.g. Stenflo 1980). This is in contrast with the degree of polarization $p=0$ resulting from a resonance transition between $1S_{1/2}$ and $2P_{1/2}$ and $p=3/7$ obtained for the $1S_{1/2}$ and $2P_{3/2}$ transition \\markcite{lee94b} (Lee et al. 1994). In a moderately thick and static medium a negligibly polarized flux is expected because the photons are locally scattered many times and get isotropized before they escape to the observer \\markcite{lee94b} (e.g. Lee 1994). However, in a very thick medium, the escape is achieved by a large number of local resonant scatterings followed by a small number of scatterings in the damping wing. Hence, in the wing regime the spatial diffusion becomes important and the radiation field may get anisotropic depending on the scattering geometry. Therefore, the emergent line photons may get polarized and also anisotropic kinematics introduced in the medium can enhance the polarization. In this {\\it Letter}, we compute the polarized flux of the emergent Ly$\\alpha$ from an optically thick and expanding slab. This result is applied to a hemi-spherical shell that is expected in various systems including primeval galaxies exhibiting P-Cygni profiles. ", "conclusions": "It seems a general consensus that the P-Cygni type Ly$\\alpha$ emissions are originated from expanding envelopes of H~II regions, which are indicative of the massive star formation. These are often obscured by dust lanes or thick molecular disks \\markcite{ich94, sco98} (Ichikawa et al. 1994, Scoville et al. 1998). In this {\\it Letter} we computed the polarization of the Ly$\\alpha$ photons that are transferred through an optically thick and expanding neutral hydrogen layer. Anisotropic expansion and high column density are coupled to enhance scatterings in the damping wing into the direction corresponding to the largest velocity gradient, which results in highly polarized emergent flux. In particular, in a spherical shell with column density $\\sim 10^{20}~\\cm^{-2}$ and expansion velocity $\\sim 100~\\kms$ we find that the averaged degree of polarization of the emergent Ly$\\alpha$ line photons reach as high as 0.05 when $\\mu = 0.5$. There are three interesting classes of primeval objects showing P-Cygni type profiles; the DLA candidate galaxies including DLA~2233+131 \\markcite{djor96, lu96, lu97} (Djorgovski et al. 1996, Lu et al. 1996, 1997) and DLA~2247-021 \\markcite{djor97} (Djorgovski 1997), the Lyman break galaxies at $310\\pm 3$) and the total luminosity lower by a factor of $\\approx 6$. There is a velocity gradient of $\\approx 300$ km s$^{-1}$ in the line emission across the continuum knot, in the sense that the side nearest the radio galaxy is blueshifted, and that on the far side is, within errors of $\\approx 100\\, {\\rm km\\, s^{-1}}$, at rest with respect to the starlight in `c'. There is also a faint tail of yet more highly blueshifted emission (up to $\\approx 600\\, {\\rm km\\, s^{-1}}$) on the side nearest the radio galaxy, pointing towards the radio galaxy. Note that there is no sign of emission lines in Fig.\\ 4, which used a narrow extraction about the continuum peak. In contrast, in Fig.\\ 5 the extended emission-line flux from `c' is clearly visible and is quite strong when integrated over the entire emission region (Table 1). \\setlength{\\unitlength}{1mm} \\begin{figure*} \\begin{picture}(150,60) \\put(-25,-200){\\special{psfile=linefig.ps}} \\end{picture} \\caption{\\small{2-D spectrum of 3C441 with the slit aligned along the axis joining the radio galaxy and `c'. Left: the [O{\\sc ii}] emission line with wavelength on the horizontal axis and distance along the slit vertically. The radio galaxy is to the top and component `c' below it. Right: the [O{\\sc iii}]5007 emission line. The figures are $11.2\\, {\\rm nm} \\times 30\\, {\\rm arcsec}$ in size (11.2 nm $\\approx$ 5280 km s$^{-1}$ close to [O{\\sc ii}] and $\\approx$ 3930 km s$^{-1}$ close to [O{\\sc iii}]). The contour levels are spaced at intervals of $2\\times 10^{-21} {\\rm Wm^{2}nm^{-1}}$ per pixel, starting from $2\\times 10^{-21} {\\rm Wm^{2}nm^{-1}}$, each pixel was $0.28 {\\rm nm} \\times 0.33 {\\rm arcsec}$ in size.}} \\end{figure*} \\subsection{The radio structure} The asymmetric radio structure of 3C441 is interesting in the context of the models to explain the aligned emission. The asymmetry in jet brightness either side of the nucleus is very pronounced, and can be interpreted either as an asymmetry produced by Doppler boosting, or in terms of 3C441 having a radio structure transitional between FRI and FRII, with one side FRII-like and the other more FRI-like. The lack of a radio central component, commonly seen in radio galaxies with Doppler boosted one-sided jets (e.g. 3C22, Rawlings et al.\\ 1994), argues strongly in favour of the latter explanation, and indeed the flaring of the jet just to the NW of the host galaxy is reminiscent of the ``Mach disk'' structure seen in the M87 jet (Owen, Hardee \\& Cornwell 1989). \\subsection{The radio galaxy} The spectrum of `a', the host galaxy of the radio source is shown in Fig.\\ 2. It is a typical moderately-high ionisation narrow-line radio galaxy spectrum, with strong emission lines superposed on a stellar spectrum dominated by an old stellar population with a strong 4000\\AA$\\;$break. Close inspection of the images shows, however, that even in this case, where the integrated light is dominated by old stellar populations there is evidence for morphological peculiarity. In particular there is a distinct blue aligned component to the south east of the host galaxy peak (Fig.\\ 6). Whether this represents a spiral arm, a merger remnant or some form of radio source-induced aligned component is unclear. As its separation from the radio galaxy is only about 0.5-arcsec, it is hard to tell from the spectra whether it is line or continuum dominated, but the more diffuse material extending $\\approx 2$-arcsec to the south of the radio galaxy is definitely continuum dominated. Although much weaker relative to the smooth underlying host galaxy in the F785LP image, the aligned component is nevertheless visible, along with some more diffuse aligned emission on the other side of the peak, to the northwest. In Fig.\\ 7, the position angle of the host galaxy (derived from the second moments of the flux distribution) is plotted for various isophotes and apertures. This shows two interesting aspects of the alignment. First, the alignment persists into the $K$-band, suggesting the aligned light is fairly red. Second, in the {\\em HST} images, there is evidence for isophotal twisting as the aperture size is increased to include the inner aligned components highlighted in Fig.\\ 6 (PA $\\sim 150$ deg), and later the low surface brightness emission round the host at PA $\\sim 0$, seen best in Fig.\\ 1. The low-surface brightness material to the northwest may be mostly line emission though, as the emission-line image of McCarthy et al.\\ (1995) shows that the [O{\\sc ii}] emission is also roughly aligned along PA 0. \\setlength{\\unitlength}{1pt} \\begin{figure} \\begin{picture}(200,100) \\put(-240,-350) {\\special{psfile=3c441rg_555.ps}} \\put(-120,-350){\\special{psfile=3c441rg_785.ps}} \\put(0,80){(a)} \\put(120,82){(b)} \\end{picture} \\caption{Close-up of the host galaxy (`a') greyscaled so as to show the aligned components near the nucleus: (a) F555W image; (b) F785LP image. Both images are 10 arcsec square and have been smoothed with a $\\sigma = 0.05$ arcsec gaussian.} \\end{figure} \\setlength{\\unitlength}{1pt} \\begin{figure*} \\begin{picture}(500,200) \\put(-165,-250){\\special{psfile=kbandten_bm.ps vscale=80 hscale=80}} \\put(15,-250){\\special{psfile=ibandten_bm.ps vscale=80 hscale=80}} \\put(195,-250){\\special{psfile=vbandten_bm.ps vscale=80 hscale=80}} \\end{picture} \\caption{Position angle as a function of aperture size and isophotal cutoff for the host galaxy of 3C441. The isophotal cutoff level is in units of the sky noise, and the aperture radii (indicated by circles of increasing size) are 0.6, 0.9, 1.1, 1.7 and 2.4 arcsec in (a), and 0.3, 0.4, 0.6, 0.8, 1.2, 1.7, and 2.4 arcsec in (b) and (c). The radio source PA (defined as the PA of the line joining the hotspots) is indicated by a dotted line.} \\end{figure*} \\subsection{The relationship of the radio and optical components} The relative astrometry was performed using the results of the APM scans of the Palomar Sky Survey plates. The positions of four stars were used to align the corners of the radio map with the optical images. The uncertainty in the overlay from the scatter in the fit was $\\approx 0.3$ arcsec. This astrometry places the host galaxy (`a' in Fig.\\ 1a) just to the SE of the first appearance of the northern radio jet. Component `d' is apparently aligned along the jet direction, although positioned to the side of it. The point of deflection of the jet is approximately coincident with the peak of the [O{\\sc ii}]372.7 emission, about 3-arcsec SE of the peak of the continuum knot `c'. In the F555W image (Fig.\\ 1a), there is an arc of emission just to the north of the north-west radio hotspot and apparently centered on `c'. Fig.\\ 1b shows the F785LP image; here there is a halo of diffuse emission around `c'. By subtracting a spectrum centered on the continuum knot in an aperture 1.7-arcsec wide from the total spectrum of the aligned component `c' (in a 6.8-arcsec wide aperture), we have been able to estimate the line contribution to the continuum magnitudes measured for the nebular region surrounding `c'. In both filters this is $\\approx 10$ per cent overall, but in the regions of brightest line emission in the interaction region to the south of `c' our spectrum suggests that the line contribution of [O{\\sc ii}] to the F555W flux rises to dominate the overall flux in the filter, consistent with the arc of emission seen in the F555W region just above the radio hotspot consisting entirely of line emission. The continuum emission present in this extraction is bluer than the emission from the knot. The linear object `d' (Fig.\\ 8) is reminiscent of the aligned component in 3C34 (Best, Longair \\& R\\\"{o}ttgering 1997a). Like the object in 3C34, it has no emission lines visible either in our spectra or in the narrow-band image of McCarthy et al.\\ (1994), but is well aligned with the radio structure and lies within the radio lobe. Its optical--near infrared colours are bluer than `c' ($K=20.8; J>22.7; m_{785}=23.3; m_{555}=25.1$ in 3-arcsec diameter apertures, where $m_{785}$ and $m_{555}$ are AB magnitudes in the two {\\em HST} images), but show a break between the 785LP and 555W filters, consistent with a 4000$\\;$\\AA$\\;$break at the redshift of the radio source. \\setlength{\\unitlength}{1pt} \\begin{figure} \\begin{picture}(200,100) \\put(-548,-740){\\special{psfile=3c441d_555.ps hscale=200 vscale=200}} \\put(-430,-740){\\special{psfile=3c441d_785.ps hscale=200 vscale=200}} \\put(0,80){(a)} \\put(120,80){(b)} \\end{picture} \\caption{Close-up of component `d': (a) F555W image; (b) F785LP image. Both images are 5 arcsec square and have been smoothed with a $\\sigma = 0.1$ arcsec gaussian.} \\end{figure} \\subsection{Other companion objects} There are several objects around the radio galaxy which appear to be members of a group or cluster around it. 3C441 is in a sample of radio galaxies whose clustering properties we are currently evaluating, and a formal estimate of the clustering amplitude, $B_{\\rm gq}$, for this radio source will be presented in Wold et al.\\ (in preparation). For the purposes of this paper, we simply compared the counts of galaxies with magnitudes $m_1$ to $m_1+3$ (where $m_1$ is the magnitude of the radio source host galaxy) in the object frame of the F785LP image (the WF3 CCD) with the average of those in the two side frames (WF2 and WF4). This revealed an excess of $11.5 \\pm 7.1$ galaxies, indicating the possible presence of a cluster, but not at a high confidence level. Clearly though this is likely to be an underestimate of the cluster richness as many cluster members may be outside the restricted field of the CCD, and present on the other frames, increasing the estimate of the background count." }, "9803/astro-ph9803221_arXiv.txt": { "abstract": "We modified the Press-Schechter (PS) formalism and then analytically derived a constrained mass distribution function $n(M|\\varphi)$ for the regions having some specified value of the primordial gravitational potential, $\\varphi$. The resulting modified PS theory predicts that gravitationally bound clumps with masses corresponding to rich clusters are significantly biased toward the regions of negative primordial potential - the troughs of the potential. The prediction is quantitative, depending on the mass and the depth of the troughs, which can be tested in large N-body simulations. As an illustration of the magnitude of the effect we calculate the constrained mass function for the CDM model with $\\Gamma = \\Omega h = 0.25$ normalized to $\\sigma_{8} = 1$. In particular, we show that the probability of finding a clump of mass $10^{14} - 10^{15}h^{-1}M_{\\odot}$ in the region of negative initial potential is $1.3 - 3$ times greater (depending on the mass) than that in the region of positive initial potential. The scale of the potential fluctuations $R_{\\varphi}=\\sqrt{3} \\sigma_{\\varphi}/\\sigma_{\\varphi'}$ is shown to be $\\approx 120 h^{-1}{\\rm Mpc}$ for the spectrum in question. The rms mass density contrast on this scale is only about $\\sigma _{\\delta}(R_{\\varphi}) \\approx 0.03$. Assuming that the modified PS theory is statistically correct, we conclude that clusters are significantly biased ($b \\ge 10$, $b$ is a bias factor defined by $\\Delta n_{cl}/ n_{cl} =b \\Delta \\rho_m/\\rho_m$) toward the regions having negative initial potential. ", "introduction": "Assuming the standard hierarchical model of the structure formation from Gaussian fluctuations due to gravitational instability, we study the effect of primordial gravitational potential fluctuations on massive objects such as galaxy clusters and perhaps superclusters, i.e., clusters of clusters (\\cite{bah-son84}). We employ the PS formalism as a tool and modify it for this study. Some effect of the primordial gravitational potential upon the structure formation has been already noted. \\cite{kof-sha88} have noticed that the adhesion approximation predicts that the formation of voids is associated with positive peaks of the primordial gravitational potential. Sahni, Sathyaprakash, \\& Shandarin (1994) studied the effect and measured a significant correlation between the sizes of voids and the value of primordial gravitational potential in numerical simulations of the adhesion model. By investigating the evolution of correlation between the potential and the density perturbations, Buryak, Demianski, $\\&$ Doroshkevich (1992) showed that the formation of super large scale structures is mainly determined by the spatial distribution of the gravitational potential. Recently, Madsen et al. (1997) have demonstrated by N-body simulations that the under dense and the over dense regions are closely linked to the regions with the positive and the negative gravitational potential respectively. Thus, given all these results showing the important role of the primordial gravitational potential in the structure formation, it would be interesting to calculate the effect of the primordial potential upon the mass distribution function. The mass distribution function $n(M)$ is defined such that $n(M)dM$ is the comoving number density of gravitationally bound objects in the mass range $(M,M + dM)$. The standard Press-Schechter (hereafter, PS) formalism provides an effective tool to evaluate $n(M)$ in spite of various criticism on it (see \\cite{mon98}), and is widely used in cosmology (e.g. \\cite{gro-etal97}; \\cite{kit-sut97}; \\cite{bah-fan98}; \\cite{rob-gaw-sil98}; \\cite{wan-ste98}). Also, \\cite{lee-sh98} have shown by applying the dynamics based on the Zel'dovich approximation to the PS formalism that it is very robust with respect to the underlying dynamics. The following two equations represent the essence of the PS formalism (\\cite{pre-sch74}): \\begin{equation} n(M) = \\frac{\\bar{\\rho}}{M}\\bigg{|} \\frac{dF}{dM}\\bigg{|} , \\end{equation} \\begin{equation} F(M) = \\int^{\\infty}_{\\delta_{c}}\\! p(\\delta) d\\delta. \\end{equation} Here $p(\\delta)$ is the probability density distribution of the linearly extrapolated density contrast $\\delta$ smoothed on a comoving filtering scale $R$ which is related to the mass by $M=M(R)=\\alpha{\\bar\\rho}R^3$. The proportionality constant $\\alpha$ is either determined by the shape of the smoothing window function or sometimes is used as a free parameter in order to provide a better agreement with numerical results. In the case of a sharp k-space filter which is actually consistent with the PS formalism (see \\cite{pea-hea90}), the filtering scale $k_c = 2\\pi/R$ in k-space and mass are related as $M = 6\\pi^{2}\\bar{\\rho}k_c^{-3}$. The density threshold value $\\delta_c$ for collapse was originally given as $\\delta_{c}\\approx 1.69$ according to the Top Hat spherical model. However, it has been shown that the lowered value of $\\delta_c$ in the range from $1.3-1.6$ gives a better fit in N-body simulations, which depends on the the initial spectrum and the type of the filter (e.g., \\cite{gro-etal97}). In this Letter we investigate and show how much the primordial gravitational potential $\\varphi$ affects the mass distribution function of galaxy cluster. Modifying the PS formalism, we derive a constrained mass distribution function $n(M|\\varphi$) defined as the comoving number densities of clumps of mass $M$ in the regions where the primordial gravitational potential fluctuation satisfies some specified conditions. The Cold Dark Matter model (CDM) with $\\Gamma = \\Omega h = 0.25$ and $\\sigma_8=1$ is used to demonstrate the significance of the effect. ", "conclusions": "The PS formalism has been proved to be a simple but very effective tool widely used for constraining cosmological models. We have modified it by considering the dependence of mass function on the initial perturbation of gravitational potential. The resulting modified PS theory predicts that the clumps with masses greater than roughly $10^{14}h^{-1}M_{\\odot}$ have a noticeable tendency to form in the troughs of the primordial gravitational potential (the regions where the primordial potential fluctuations were negative). This quantitative prediction can be tested in large N-body simulations. Regardless of the outcome it will shed light on the PS formalism; if our prediction is confirmed, it will show a new potency of the PS technique. Otherwise a new limitation to the formalism will be established. Assuming that the prediction is correct at least qualitatively, \\footnote{N-body simulations (e.g., Madsen et al. 1997) and the adhesion model (Sahni et al. 1994) have already visually demonstrated this bias effect of the gravitational potential.} we would like to discuss some of its obvious consequences. The scale of the initial potential \\begin{equation} R_{\\varphi} = \\sqrt{3} \\sigma_{\\varphi}/ \\sigma_{\\varphi'} =\\sqrt{3\\frac{\\int^{\\infty}_{k_{l}}\\! dk k^{-2}P(k)} {\\int^{\\infty}_{0}\\! dk P(k)}} \\approx 120 h^{-1} {\\rm Mpc} \\end{equation} does not depend on any ad hoc scale; the dependence on $k_l$ is exremely weak ($\\propto \\sqrt{ln{(1/k_l)}}$ for the Harrison-Zel'dovich spectra assumed here). It is, perhaps, worth mentioning that the scale of the potential is also practically independent of the smoothing scale unless it exceeds the value of a few tens of $h^{-1}{\\rm Mpc}$. The density scale $R_{\\delta_{k_c}}$ is determined by the scale of the smoothing window function $k_c$ that has only one ``natural'' scale corresponding nonlinearity $k_c=k_{nl}$. For the model in question the scale of the primordial potential is found to be $R_{\\varphi} \\approx 120 h^{-1} {\\rm Mpc}$. The scale of the density contrast field reaches this value $R_{\\delta} = \\sqrt{3} \\sigma_{\\delta}/\\sigma_{\\delta'} \\approx 120 h^{-1} {\\rm Mpc}$ only after it is smoothed on $k_c \\approx 0.017 h {\\rm Mpc^{-1}}$. The corresponding density variance on this scale is $\\sigma_{\\delta}(0.017 h {\\rm Mpc^{-1}}) \\approx 0.03$. On the other hand, the number of clumps with masses $10^{14} - 10^{15} h^{-1} M_{\\odot}$ can easily be 30\\% greater in the troughs of the potential than the mean density $n(>M) = 0.5[n(>M|\\varphi<0)+ n(>M|\\varphi>0)]$ (see Fig. 1). Thus, the bias factor $b$ (defined by the relation $\\Delta n_{cl}/ n_{cl} = b \\Delta \\rho_m/\\rho_m$) reaches at least $10$ on the scale about $120 h^{-1} {\\rm Mpc}$. Qualitatively the bias phenomenon can be explained as follows. The initial density contrast is proportional to the Laplacian of the initial potential ($\\delta \\propto \\nabla^2 \\varphi$). Therefore the two fields are cross-correlated: the positive peaks of $\\delta$ are more likely to be found in the troughs of the potential where it is negative. The correlation is not very strong (for $k_c = 0.25 h {\\rm Mpc^{-1}}$ corresponding to $\\sigma_{\\delta} = 1$ the crosscorrelation coefficient $\\kappa = \\sigma_v^2(0.25 h {\\rm Mpc^{-1}})/ \\sigma_{\\varphi} \\sigma_{\\delta}(0.25 h {\\rm Mpc^{-1}}) \\approx 0.12$). But the clusters are extreme objects corresponding to the tail of the mass function, and thus very sensitive to the environment. That is why the clusters put one of the strongest constraints on cosmological models (\\cite{kly-rhe94}, \\cite{bo-my96}, \\cite{fan-etal97}, \\cite{bah-fan98}). Incorporating the motion of mass into dynamics can only increase the bias effect due to the nonlinear effects although they are quite small on the scale in question. But, the point is not in the magnitude of the nonlinear effects but rather in their sign. On the scale of the potential the mass moves from the peaks of the potential to the troughs. Using the Zel'dovich approximation one can easily estimate the rms displacement of the mass on the scale of the potential (\\cite{sh93}): \\begin{equation} d_{rms} = \\sqrt{{{\\int_0^{0.017h}P(k) dk} \\over {\\int_0^{0.25h}P(k)k^2 dk}}} \\approx 3 h^{-1} {\\rm Mpc}. \\end{equation} It is relatively small compared to the scale of the potential but coherent on the scale of the potential field, and therefore it can only enhance the bias effect. Another nonlinear effect is related to the rate of growth of perturbations. For the perturbations on the scale of a few Mpc the potential troughs/peaks may be viewed as patches with slower/faster expansion rate that corresponds to the increase/decrease of the rate of growth of small-scale perturbations. Similarly, the bias is enhanced in the redshift space because the velocity field is directed toward the troughs and away from peaks of the potential. Both effects can increase the bias by about 5\\% depending on the initial spectrum. Another way of calculating the constrained mass function would be using the peak-background split technique suggested by \\cite{kai84} to explain the enhanced correlation function of reach clusters. Obviously, the initial potential resembles the smoothed initial density field if the filter has a sufficiently large scale, but the former is never identical to the latter. The potential itself can be viewed as a smoothed density field with a very soft scale-free filter $W(k) \\propto k^{-2}$. Typically the density field is filtered with much harder filters (e.g. top-hat, Gaussian, or sharp $k$-space filters), that impose the scale which is an ad hoc parameter. The magnitude of the bias in our approach is determined by the crosscorrelation of the density contrast smoothed at the scale ($k_c$) of nonlinearity ($\\sigma_{\\delta_{k_c}}=1$) with the initial potential that does not have any ad hoc parameters. Probably, the value of the crosscorrelation coefficient determines the bias in the peak-split approach as well. The crosscorrelation of the density field $\\delta_{k_c}$ smoothed on the scale of nonlinearity ($k_c = 0.25 h {\\rm Mpc^{-1}}$) with the field $\\delta_{k_{\\varphi}}$ smoothed on the scale of the potential ($k_{\\varphi} = 0.017 h {\\rm Mpc^{-1}}$) is about $4$ times weaker than the correlation of $\\delta_{k_c}$ with the initial potential $\\varphi$. Thus, we expect that the bias of galaxy clusters on such large scales as the scale of the initial potential ($\\approx 120 h^{-1} {\\rm Mpc}$) is stronger toward the troughs of the potential fluctuations rather than to the peaks of the density fluctuations $\\delta_{k_c}$ smoothed with the corresponding filter. We have not applied the split peak-background approach because it is not clear how to avoid arbitrarines in choosing the scale that splits the density into small-scale peaks and large-scale background field. This question requires a separate study. Applying this effect to observations one has to take into account the following issues. The gravitational potential does not evolve much on large scales especially in the Einstein-de Sitter universe (Kofman \\& Shandarin 1988; \\cite{pau-mel95}; \\cite{mel-etal96}). Therefore, the potential at present is very similar to the primordial one on scales much greater than the scale of nonlinearity. A simple explanation to this in the frame of the standard scenario of the structure formation is due to the fact that the mass has been displaced by the distance about $10 h^{-1}{\\rm Mpc}$ (\\cite{sh93}). Therefore, the potential on scales greater than, say, $30 h^{-1}{\\rm Mpc}$ has been changed very little. Clusters can be used as {\\it statistical} tracers of the potential. In addressing this question it is worth noting that the shot noise is an important factor since clusters are rare objects. Using the observational mass function (Bahcall \\& Cen 1993) one can estimate that an average spherical patch of the radius $\\approx 60 h^{-1} {\\rm Mpc}$ contains about $30$ clusters with the masses greater than $10^{14} h^{-1} M_{\\odot}$. Thus, the shot noise is about 18\\% on this scale which is comparable with the bias itself (see Fig. 1, the bottom panel). However, Fig. 2 suggests that the most massive clusters [$M>10^{15} h^{-1} M_{\\odot}$] are very likely to reside in the regions of negative potential ($P > 75\\%$) and very unlikely in the regions of high potential ($P < 5\\%$ if $\\varphi > \\sigma_{\\varphi}$). More detailed analysis will be present elsewhere. Probably, the best candidates for the markers of the troughs in the field of the primordial potential fluctuations are superclusters (defined as clusters of clusters) (\\cite{bah-son84}) especially with highest density enhancements (Shapley supercluster) and the giant geometrical patterns in the cluster distribution (\\cite{tul-etal92})." }, "9803/hep-ph9803378_arXiv.txt": { "abstract": "The scattering of solar neutrinos on electrons is sensitive to the neutrino magnetic moments through an interference of electromagnetic and weak amplitudes in the cross section. We show that future low-energy solar neutrino experiments with good angular resolution can be sensitive to the resulting azimuthal asymmetries in event number and should provide useful information on non-standard neutrino properties such as magnetic moments. We compare asymmetries expected at Hellaz (mainly pp neutrinos) with those at the Kamiokande and Super-Kamiokande experiments (Boron neutrinos), both for the case of Dirac and Majorana neutrinos and discuss the advantages of {\\sl low energy experiments}. Potentially interesting information on the solar magnetic fields may be accessible. ", "introduction": "Most non-standard properties of neutrinos arise from non-zero masses \\cite{fae,revnu}. Among these electro-magnetic dipole moments play an important role \\cite{MoPal}. Here we are concerned with a particular effect in neutrino-electron scattering for neutrinos from the Sun which possess a Dirac magnetic moment \\cite{VVO} or transition magnetic moments \\cite{BFD} in the case of Majorana neutrinos. The latter is especially interesting first of all because it is more fundamental theoretically, and because Majorana neutrinos are the ones which arise in most extensions of the Standard Model. Moreover, the effects of Majorana transition moments can be resonantly enhanced when neutrinos propagate in media \\cite{RFSP} such as the Sun, providing one of the attractive solutions to the solar neutrino problem \\cite{akhmedov97}. Another practical advantage in favour of Majorana transition moments is that, in contrast to Dirac-type magnetic moments, these are substantially less stringently constrained by astrophysics \\cite{Raffelt}. For {\\sl pure left-handed neutrinos} the weak interaction and the electro-magnetic interaction amplitudes on electrons do not interfere, since the weak interaction preserves neutrino helicity while the electro-magnetic does not. As a result the cross section depends quadratically on $\\mu_\\nu$. However, if there exists a process capable of converting part of the initially fully polarized $\\nu_e$'s, then an {\\sl interference term} arises, proportional to $\\mu_\\nu$, as pointed out e.g. in ref. \\cite{Barbieri}. This term depends on the angle between the component of the neutrino spin transverse to its momentum and the momentum of the outgoing recoil electron. Therefore the event count rates expected in an experiment would exhibit an {\\sl asymmetry} with respect to the above defined angle. Such asymmetry would not show up in earth-bound laboratory experiments even with stronger magnetic fields, since the helicity-flip could be caused only by the presence of a neutrino mass and is therefore negligible \\cite{grimus}. However, in the solar convective zone one may find a magnetic field extended over a tenth or so of the solar radius and, most importantly, the neutrino depolarization could be resonant in the Sun. Even if the Sun possesses only a relatively modest large-scale magnetic field $B_{\\perp} \\sim 10^4$ G in the convective region ($L\\sim L_{conv} \\simeq 3\\times 10^{10}$ cm), and for a neutrino magnetic moment of the order $10^{-11} \\mu_B$ such a spin-flip process may take place with sizeable rates, since in such a case one has $\\mu_{\\nu} B_{\\perp}L \\sim 1 $. Barbieri and Fiorentini considered \\cite{Barbieri} the conversions $\\nu_{eL} \\to \\nu_{eR}$ in the Sun as a result of the spin-flip by a toroidal magnetic field in the convective zone. They showed that the azimuthal asymmetry could be observable in a real time solar $^8$B-neutrino experiment and as large as 20\\% for an electron kinetic energy threshold of $W_e=5$ MeV. They chose a fixed $\\nu_e$ survival probability $P_e=1/3$ (as suggested at that time by the Homestake experiment) and the maximal Dirac magnetic moment allowed by laboratory experiments, $\\mu_\\nu\\simeq 10^{-10}\\mu_B$. On the other hand, Vogel and Engel \\cite{Vogel} emphasized that if an asymmetry in the scattering of solar neutrinos exists, recoil electrons will be emitted copiously along the direction of the neutrino polarization in the plane orthogonal to the neutrino momentum. They calculated the asymmetry expected for solar $^8$B neutrinos with $\\mu_\\nu=10^{-10} \\mu_B$ and concluded that it would be difficult to detect because of the poor angular resolution of the experiments. Moreover, as we will see later, both \\cite{Barbieri} and \\cite{Vogel} {\\sl overestimated} the asymmetry. Thus their calculations are not accurate. In this paper we correct results for the asymmetry in the case presented by \\cite{Barbieri} and \\cite{Vogel} for high energy $^8$B neutrinos. In addition we compare them with the expected asymmetry in the case of a Majorana transition magnetic moment of the same magnitude. More importantly, we show the sensitivity of planned solar neutrino experiments in the {\\sl low energy region} ($\\omega \\lsim 1$ MeV) to the azimuthal asymmetries that are expected in the recoil electron event rates, arising from the above electro-weak interference term. We calculate the asymmetry for the low-energy $pp$-neutrinos fixing the survival probability at $P_e=0.5$. This gives the maximum expected asymmetry and seems phenomenologically reasonable in order to convert the initial solar $\\nu_{eL}$'s via the Resonant Spin-Flavour Precession (RSFP) scenario. In particular we calculate the asymmetry that could be observed in the azimuthal distribution of events in an experiment like the proposed Hellaz \\cite{Hellaz}, sensitive to the fundamental $pp$ neutrinos from the Sun. The Multi-Wire-Chamber in Hellaz should measure both the recoil electron energy $T$ and the recoil electron scattering angle $\\theta$ with good precision. Moreover Hellaz should be sensitive to the azimuthal angle $\\phi$, measuring the number of events in $\\phi$--bins. We discuss the sensitivity of the Hellaz experiment for probing $\\mu_\\nu$ and compare it with planned accelerator experiments. In particular there are very interesting new projects, such as the future ITEP-Minnesota experiment, where they plan to search $\\mu_{\\nu}/\\mu_B$ down to $3\\cdot 10^{-11}$ with reactor anti-neutrinos \\cite{Voloshin}, and the LAMA experiment, which will use a powerful isotope neutrino source \\cite{Bernabei}. Finally, we also refine our calculations of the azimuthal asymmetry expected for $pp$ neutrinos at Hellaz using a realistic energy-dependent conversion probability $P_e$ based on a simple model for resonant spin flip conversions in the Sun. ", "conclusions": "Measuring azimuthal asymmetries in future low-energy solar neutrino-electron scattering experiments with good angular resolution should be a feasible and illuminating task. Such asymmetries should provide useful information on non-standard neutrino properties such as magnetic moments, as well as on solar magnetic fields. The effect follows from an interference of electro-magnetic and weak amplitudes in the cross section. We have seen that low-energy experiments such as Hellaz (sensitive mainly to pp neutrinos) should provide a much better means for the study of azimuthal asymmetries than accessible at the Kamiokande or Super-Kamiokande experiments (sensitive to Boron neutrinos). For equal values of the magnetic moments, the expected asymmetries are larger for Dirac neutrinos than for Majorana neutrino transition moments. However, the Dirac neutrino case is probably less likely, as there is no resonant conversion in the Sun. One exception would be the case of Dirac neutrinos in the presence of twisting magnetic fields \\cite{akhpetsmi}. However, although in this case resonant conversions in matter can take place one expects (as mentioned in section 3) a washing out of the asymmetry effect due to the changing magnetic field direction. Therefore the RSFP scenario remains as the most promising possibility. It is also the most interesting one theoretically, since Majorana neutrinos are more fundamental and arise in most models of particle physics beyond the Standard Model. Note that the discussion given above we have assumed $\\nu_e$ magnetic moments of the order $10^{-11}\\mu_B$ which is consistent with present laboratory experiments. Apart from possible effects in red giants, a $\\nu_e$ transition moment of $10^{-11}\\mu_B$ is compatible with astrophysical limits, given the present uncertainties in these considerations." }, "9803/hep-ph9803244_arXiv.txt": { "abstract": "{ \\setlength{\\baselineskip}{16pt} \\noindent Using the results of a high precision calculation of the solar neutrino survival probability for Earth crossing neutrinos in the case of MSW $\\nu_e \\rightarrow \\nu_s$\\ transition solution of the solar neutrino problem, we derive predictions for the one-year averaged day-night (D-N) asymmetries in the deformed recoil-e$^-$ spectrum and in the energy-integrated event rate due to the solar neutrinos, to be measured with the Super - Kamiokande detector. The asymmetries are calculated for three event samples, produced by solar $\\nu_e$\\ crossing the Earth mantle only, the core, and the mantle only + the core (the full night sample), for a large set of representative values of the MSW transition parameters $\\Delta m^2$ and $\\sin^2 2\\theta_V$ from the ``conservative'' $\\nu_e \\rightarrow \\nu_s$ solution regions, obtained by taking into account the possible uncertainties in the predictions for the $^{8}$B and $^{7}$Be neutrino fluxes. The effects of the uncertainties in the value of the bulk matter density and in the chemical composition of the Earth core on the predictions for the D-N asymmetries are investigated. The dependence of the D - N effect related observables on the threshold recoil - e$^-$\\ kinetic energy, $T_{e,th}$, is studied. It is shown, in particular, that for $\\sin^2 2\\theta_V \\leq 0.030$\\ the one year average D-N - asymmetry in the sample of events due to the core-crossing neutrinos is larger than the asymmetry in the full night sample by a factor which, depending on the solution value of \\dms, can be $\\sim (3 - 4)$ ($\\dms ~< 5\\times 10^{-6}~{\\rm eV^2}$) or $\\sim (1.5 - 2.5)$ ($5\\times 10^{-6}~{\\rm eV^2}~ \\ltap ~\\dms ~\\ltap 8\\times 10^{-6}~{\\rm eV^2}$). We find, however, that at small mixing angles $\\sin^2 2 \\theta_V ~\\ltap~ 0.014$, the D-N asymmetry in the case of solar $\\nu_e \\rightarrow \\nu_s$\\ transitions is considerably smaller than if the transitions were into an active neutrino, $\\nu_e \\rightarrow \\nu_{\\mu(\\tau)}$. In particular, a precision better than 1\\% in the measurement of any of the three one year averaged D-N asymmetries considered by us would be required to test the small mixing angle nonadiabatic $\\nu_e \\rightarrow \\nu_s$ solution at $\\sin^2 2\\theta_V ~\\ltap ~0.01$. For $0.0075~ \\ltap~ \\sin^2 2\\theta_V ~\\leq~ 0.03$ the magnitude of the D-N asymmetry in the sample of events due to the core-crossing neutrinos is very sensitive to the value of the electron number fraction in the Earth core, $Y_e(core)$: a change of $Y_e(core)$ from the standard value of 0.467 to the conservative upper limit of 0.50 can lead to an increase of the indicated asymmetry by a factor of $\\sim (3 - 4)$. Iso - (D-N) asymmetry contours in the $\\Delta m^2 - \\sin^2 2 \\theta_V$\\ plane for the Super-Kamiokande detector are derived in the region $\\sin^2 2\\theta_V \\geq 10^{-4}$ for the three event samples studied for $T_{e,th} = 5~{\\rm MeV ~and~7.5~MeV}$, and in the case of the samples due to the core crossing and (only mantle crossing + core crossing) neutrinos - for $Y_e(core) = 0.467~{\\rm and~} 0.50$. The possibility to discriminate between the $\\nu_e \\rightarrow \\nu_s$ and $\\nu_e \\rightarrow \\nu_{\\mu(\\tau)}$ solutions of the solar neutrino problem by performing high precision D-N asymmetry measurements is also discussed. } \\eec \\newpage ", "introduction": "\\indent In the present article we continue the systematic study of the \\daynight\\ effect for the \\SK\\ detector began in refs. \\cite{ArticleI}\\ and \\cite{ArticleII}. Assuming that the solar neutrinos undergo two - neutrino MSW $\\nue \\rightarrow \\numt$\\ transitions in the Sun, and that these transitions are at the origin of the solar neutrino deficit, we have performed in \\cite{ArticleII}\\ a high - precision calculation of the one - year averaged solar \\nue\\ survival probability for Earth crossing neutrinos, $\\PTot(\\nue\\rightarrow\\nue)$, reaching the \\SK\\ detector. The probability $\\PTot(\\nue\\rightarrow\\nue)$\\ was calculated by using, in particular, the elliptical orbit approximation (EOA) to describe the movement of the Earth around the Sun. Results for $\\PTot(\\nue\\rightarrow\\nue)$\\ as a function of $\\Enu/\\dms$, $\\Enu$\\ and \\dms\\ being the neutrino energy and the neutrino mass squared difference, have been obtained for neutrinos crossing the Earth mantle only, the core, the inner 2/3 of the core and the mantle + core (full night) for a large representative set of values of \\SdTvS, where $\\theta_V$\\ is the neutrino mixing angle in vacuum, from the ''conservative'' MSW solution region in the \\dms\\ - \\SdTvS\\ plane, derived by taking into account the possible uncertainties in the fluxes of \\BHt\\ and \\BeSv\\ neutrinos (see, e.g., ref. \\cite{SPnu96,KPUNPUB96}; for earlier studies see ref. \\cite{KS94}). We have found in \\cite{ArticleII}, in particular, that for $\\SdTvS \\leq 0.013$\\ the one - year averaged \\daynight\\ asymmetry \\footnote{A rather complete list of references on the \\daynight\\ effect is given in ref. \\cite{ArticleI}; for relatively recent discussions of the effect see, e.g., \\cite{ Hata:Langacker:1994Earth, Baltz:Weneser:1994, Gelb:Kwong:Rosen:1996, Lisi:Montanino:1997, Bahcall:Krastev:1997, Hata:1997} } in the probability $\\PTot(\\nue\\rightarrow\\nue)$\\ for neutrinos crossing the Earth core can be larger than the asymmetry in the probability for (only mantle crossing + core crossing) neutrinos by a factor of up to six. The enhancement is even larger for neutrinos crossing the inner 2/3 of the core. We have also pointed out to certain subtleties in the calculation of the time averaged \\nue\\ survival probability $\\PTot(\\nue\\rightarrow\\nue)$\\ for neutrinos crossing the Earth, which become especially important when $\\PTot(\\nue\\rightarrow\\nue)$\\ is computed, for instance, for the core crossing neutrinos only~ \\footnote{For further details concerning the technical aspects of the calculations see ref. \\cite{ArticleII} as well as ref. \\cite{NU4DN}.}. The results obtained in \\cite{ArticleII}\\ were used in \\cite{ArticleI}\\ to investigate the \\daynight\\ asymmetries in the spectrum of the recoil electrons from the reaction $\\nu + e^- \\rightarrow \\nu + e^-$\\ caused by the \\BHt\\ neutrinos and in the energy-integrated event rate, to be measured by the \\SK\\ experiment. We have computed in \\cite{SKDNII:spectrum}\\ the \\daynight\\ asymmetry in the recoil-$e^-$\\ spectrum for the same large set of representative values of \\dms\\ and \\SdTvS\\ from the ``conservative'' MSW solution region for which results in \\cite{ArticleII}\\ have been presented. The \\daynight\\ asymmetry in the $e^-$-spectrum was found for neutrinos crossing the Earth mantle only, the core and the mantle + core. In \\cite{ArticleI} we have included only 12 representative plots showing the magnitude of the \\daynight\\ asymmetry in the recoil-e$^{-}$ spectrum to be expected in the case of the two - neutrino MSW $\\nue \\rightarrow \\numt$\\ transition solution of the solar neutrino problem. The spectrum asymmetry for the sample of events due to core crossing neutrinos only was found to be strongly enhanced for $\\SdTvS~\\lsim~0.013$\\ with respect to the analogous asymmetries for the mantle and for the (only mantle crossing + core crossing) neutrinos. We presented in \\cite{ArticleI} also detailed results for the one-year averaged \\daynight\\ asymmetry in the \\SK\\ signal for the indicated three samples of events. We have found that indeed for $\\sin^22\\theta_{V}\\leq 0.013$\\ the asymmetry in the sample corresponding to core crossing neutrinos can be larger than the asymmetry in the sample produced by only mantle crossing or by (only mantle crossing + core crossing ) neutrinos by a factor of up to six. We have investigate in \\cite{ArticleI} the dependence of the D-N asymmetries in the three event samples defined above on the threshold e$^{-}$ kinetic energy being used for the event selection. The effect of the uncertainties in the Earth matter density and electron number density distributions on the predicted values of the D-N asymmetries were studied as well. We derived also iso - (\\daynight) asymmetry contours in the region of $\\SdTvS~\\gsim~10^{-4}$\\ in the \\dms - \\SdTvS\\ plane for the signals in the \\SK\\ detector, produced by neutrinos crossing the mantle only, the core and the mantle + core (full night sample). The iso-asymmetry contours for the sample of events due to core-crossing neutrinos were obtained for two values of the fraction of electrons, $Y_e$, in the core: for $Y_e (core) = 0.467$ and 0.500 \\footnote{Results for other D-N effect related observables, as the D-N asymmetry in the zenith angle distribution of the events and in the mean recoil-e$^{-}$ energy, have been obtained in refs. \\cite{Lisi:Montanino:1997, Bahcall:Krastev:1997}, where iso - (D-N) asymmetry contour plots for the full night (i.e., mantle + core) signal for the \\SK\\ detector for one value of $Y_e(core) = 0.0467$ have been presented as well.}. The results derived in \\cite{ArticleI} confirmed the conclusion drawn in \\cite{ArticleII}\\ that the \\SK\\ detector will be able to probe not only the large mixing angle adiabatic solution region, but also an important and substantial part of the $\\SdTvS~\\lsim~0.014$\\ nonadiabatic region of the MSW $\\nue \\rightarrow \\numt$\\ transition solution of the solar neutrino problem. We have found, in particular, that in a large sub-region of the ``conservative'' nonadiabatic solution region located at $\\SdTvS~\\lsim~0.0045$, the D-N asymmetry in the sample of events due to the core-crossing neutrinos only is negative and has a value in the interval (-1\\%) - (-3\\%). In the present article we realize the same program of studies for the Super-Kamiokande detector for the alternative possibility of solar neutrinos undergoing two-neutrino matter-enhanced transitions in the Sun and in the Earth into a sterile neutrino, \\nus. As is well-known, the solar $\\nu_e$ matter-enhanced transitions into a sterile neutrino, $\\nu_e \\rightarrow \\nu_s$, provide one of the possible neutrino physics solutions of the solar neutrino problem (see, e.g., \\cite{KPNPB95,Hata:Langacker:1994Earth,KPL96,SPnu96}). The reference solution region in the $\\Delta m^2 - \\sin^22\\theta$ plane, i.e., the region obtained (at 95\\% C.L.) by using the predictions of the reference solar model of Bahcall and Pinsonneault from 1995 with heavy element diffusion \\cite{BP95} ~(BP95) for the different components of the solar neutrino flux (pp, pep, $^{7}$Be, $^{8}$B and CNO), lies within the bounds: \\vspace{-1.2cm} \\bec\\beq 2.8\\times 10^{-6}~{\\rm eV}^2 ~\\ltap ~\\Delta m^2 ~\\ltap ~7.0\\times 10^{-6}~{\\rm eV}^2,~~ \\eeq\\eec \\vspace{-1.0cm} \\bec\\beq 4.8\\times 10^{-3}~ \\ltap ~\\sin^22\\theta ~\\ltap ~1.4\\times 10^{-2}~. \\eeq\\eec \\noindent The reference solution is of the small mixing angle nonadiabatic type. A reference large mixing angle (adiabatic) solution (present in the case of $\\nu_{e} \\rightarrow \\nu_{\\mu (\\tau)}$ transitions) is practically ruled out by the solar neutrino data \\cite{KPNPB95,KPL96,SPnu96}. If we allow for possible uncertainties in the predictions for the fluxes of the $^{8}$B and $^{7}$Be neutrinos \\cite{KPL96}, the solar neutrino data is described in terms of the hypothesis of $\\nu_e \\rightarrow \\nu_s$ transitions for larger ranges of values of the parameters $\\Delta m^2$ and $\\sin^22\\theta$, belonging to the intervals: \\vspace{-1.2cm} \\bec\\beq 2.8\\times 10^{-6}~{\\rm eV}^2 ~\\ltap ~\\Delta m^2 ~\\ltap ~8.0\\times 10^{-6}~{\\rm eV}^2,~~ \\eeq\\eec \\vspace{-1.0cm} \\bec\\beq 8.0\\times 10^{-4} ~\\ltap ~\\sin^22\\theta ~\\ltap ~3.0\\times 10^{-2}~. \\eeq\\eec \\vspace{-0.4cm} \\noindent and \\vspace{-1.2cm} \\bec\\beq 5.3\\times 10^{-6}~{\\rm eV}^2 ~\\ltap ~\\Delta m^2 ~\\ltap ~1.2\\times 10^{-5}~{\\rm eV}^2,~~ \\eeq\\eec \\vspace{-1.0cm} \\bec\\beq 0.13 ~\\ltap ~\\sin^22\\theta ~\\ltap ~0.55~. \\eeq\\eec \\noindent The ``conservative'' solution regions lying within the bounds determined by eqs. (3) - (4), and eqs. (5) - (6) have been obtained by treating the $^{8}$B neutrino flux as a free parameter in the relevant analysis of the solar neutrino data, while the the $^{7}$Be neutrino flux was assumed to have a value in the interval $\\Phi_{Be} = (0.7 - 1.3)~\\Phi^{BP}_{Be}$, where $\\Phi^{BP}_{Be}$ is the flux in the reference solar model \\cite{BP95}. For values of $\\Delta m^2$ and $\\sin^22\\theta$ from the solution regions (5) and (6) the $\\nu_e \\rightarrow \\nu_s$ transitions of $^{8}$B neutrinos having energy $E_{\\nu} \\geq 5~{\\rm MeV}$ are adiabatic. However, this adiabatic solution is possible only for large values of the $^{8}$B neutrino flux \\cite{KPL96}, $\\Phi_{B} \\cong (2.5 - 5.0)~\\Phi^{BP}_{B}$, $\\Phi^{BP}_{B}$ being the reference model flux \\cite{BP95}. Such values of $\\Phi_{B}$ seem totally unrealistic from the point of view of the contemporary solar models and we consider the indicated adiabatic solution here for completeness. It should be added that in deriving the conservative solution regions represented by eqs. (3) - (4) and eqs. (5) - (6) the limit on the D-N effect derived in \\cite{Hata:Langacker:1994Earth} on the basis of the data obtained in the Kamiokande II and III experiments \\cite{KamDN} was utilized. Let us note that the preliminary result on the D-N effect from the Super-Kamiokande experiment after approximately one year (374.2 days) of data taking reads \\cite{SKSB97}: \\vspace{-0.6cm} \\bec\\beq \\bar{A}^{SK}_{D-N} \\equiv \\, \\frac{\\bar{R}^{D} - \\bar{R}^{N}} {\\bar{R}^{D} + \\bar{R}^{N}} = - 0.031 \\pm 0.024 \\pm 0.014, \\eeq\\eec \\vspace{0.2cm} \\noindent where $\\bar{A}^{SK}_{D-N}$ is the average energy integrated D-N asymmetry and $\\bar{R}^{D}$ and $\\bar{R}^{N}$ are the observed average event rates caused by the solar neutrinos during the day and during the night in the Super-Kamiokande detector in the period of data taking. The first error in eq. (7) is statistical and the second error is systematic. The data were obtained with a recoil$-e^{-}$ threshold energy $E_{e,th} = 6.5~{\\rm MeV}$. In addition of performing i) detailed high precision calculations of the D-N asymmetries in the recoil-e$^{-}$ spectrum and ii) of the energy-integrated D-N asymmetries for the three samples of events (due to core crossing, only mantle crossing and only mantle + core crossing neutrinos), and of studying iii) the effects of the recoil-e$^{-}$ energy threshold variation and iv) of the uncertainties in the chemical composition and matter density of the Earth's core on the calculated D-N effect related observables, we also analyze qualitatively the possibility to distinguish between the two solutions of the solar neutrino problem involving matter-enhance transitions of the solar $\\nu_e$ respectively into active neutrinos and into sterile neutrinos, $\\nu_e \\rightarrow \\nu_{\\mu (\\tau)}$ and $\\nu_e \\rightarrow \\nu_s$, by performing high-precision D-N asymmetry measurements. \\vspace{-0.3cm} ", "conclusions": "\\indent In the present article we have performed a rather detailed quantitative study of the D-N effect for the Super-Kamiokande detector for the solution of the solar neutrino problem involving two-neutrino matter-enhanced transitions of the solar neutrinos into a sterile neutrino, $\\nu_e \\rightarrow \\nu_s$. The one year average D-N asymmetry, \\AsymRs, has been calculated (using the high precision methods developed in refs. \\cite{ArticleI,ArticleII}) for three samples of events, \\mantle\\ (M) , \\core\\ (C) and \\night\\ (N), produced respectively by the solar neutrinos crossing the Earth mantle only, the Earth core, and by the only mantle crossing + the core crossing neutrinos (the full night sample). The asymmetry calculations require the knowledge of the one year averaged spectrum of the recoil electrons and energy-integrated even rate, produced by the solar neutrinos during the day (the \\DAY\\ sample). Results for the D-N asymmetry in the recoil-e$^{-}$ spectrum for the same three samples of events, \\AsymSs (\\Te), s=N,C,M, have also been obtained. The asymmetries have been calculated for a large representative set of values of the neutrino transition parameters \\dms\\ and $\\sin^22\\theta$ from the ``conservative'' $\\nu_e \\rightarrow \\nu_s$ transition solution regions (eqs. (3) - (6)), derived by taking into account the possible uncertainties in the predictions for the $^{8}$B and $^{7}$Be neutrino fluxes. We have investigated the dependence of the three D-N asymmetries studied, on the recoil-e$^{-}$ kinetic energy threshold $T_{e,th}$, which can be varied in the Super-Kamiokande experiment, by performing calculations of all the indicated D-N asymmetries for $T_{e,th} = 5.0~$ MeV and $T_{e,th} = 7.5~$ MeV. The effect of the estimated uncertainties in the knowledge of the bulk matter density and the chemical composition of the Earth core \\cite{Stacey:1977,PREM81,CORE} on the predictions for the D-N asymmetries, has been studied as well by deriving results for \\AsymSs (\\Te)\\ and \\AsymRs\\ both for the standard value of the electron number fraction in the core $\\Ye (core) = 0.467$ and for the estimated conservative upper limit on $\\Ye (core)$, $\\Ye (core) = 0.50$ (see \\cite{CORE} and \\cite{ArticleI}). Iso-(D-N) asymmetry contour plots for the \\night, \\core\\ and \\mantle\\ samples of events in the region $10^{-7}~{\\rm eV^2} ~\\leq ~\\dms~\\leq 10^{-4}~{\\rm eV^2}$, $10^{-4}~\\leq~ \\sin^22\\theta_V ~\\leq~ 1$, have been obtained for $T_{e,th} = 5.0~$ MeV and $T_{e,th} = 7.5~$ MeV, and for the \\night\\ and \\core\\ samples - for $\\Ye (core) = 0.467$ and $\\Ye (core) = 0.50$. The main results of this study are collected in Tables I - VII and are shown graphically in Figs. 4 - 7. We have found that, as like in the case of the $\\nue \\rightarrow \\numt$\\ solution, the division of the data collected at night into a \\core\\ and \\mantle\\ samples is a rather effective method of enhancing the D-N asymmetry at small mixing angles, $0.001~\\ltap~ \\sin^22\\theta_V ~\\ltap~ 0.03$: the asymmetry in the \\core\\ sample $|\\AsymRC|$ is larger than the asymmetry in the \\night\\ sample $|\\AsymRN|$ typically by a factor of (3 - 4) if $\\dms ~< 5\\times 10^{-6}~{\\rm eV^2}$, and by a factor of $\\sim (1.5 - 2.5)$ for $5\\times 10^{-6}~{\\rm eV^2}~ \\ltap ~\\dms ~\\ltap 8\\times 10^{-6}~{\\rm eV^2}$ (Table II). However, the enhancement is not as strong as in the case of the $\\nue \\rightarrow \\numt$\\ transition solution \\cite{ArticleI}. Moreover, in the interesting region $0.005~\\ltap~ \\sin^22\\theta_V ~\\ltap~ 0.014$ the D-N asymmetries in the \\core\\ and \\night\\ samples found for the $\\nue \\rightarrow \\nus$\\ solution, $|\\AsymRC (sterile)|$ and $|\\AsymRN (sterile)|$, are substantially smaller - at least by a factor of 4 and typically by a factor of 5 to 10, than the asymmetries corresponding to the $\\nue \\rightarrow \\numt$\\ solution, \\AsymRC (active) and \\AsymRN (active). Similar conclusion is valid for the \\mantle\\ sample asymmetries. This remarkable difference in the magnitudes of the asymmetries $|\\AsymRs (sterile)|$ and $|\\AsymRs (active)|$ in the corresponding small mixing angle solution regions is a consequence of the different roles the neutron number density distribution in the Earth $n_{n}(r)$ plays in the solar neutrino transitions in the two cases: the $\\nue \\rightarrow \\numt$\\ transitions, as is well-known, depend only on the electron number density distribution, $n_{e}(r)$, while the $\\nue \\rightarrow \\nus$\\ transitions depend on the difference ($n_{e}(r) - 0.5~n_{n}(r)$). In the Sun one has \\cite{BP95} $0.5~n_{n}(r) \\ll n_{e}(r)$ and $n_{n}(r)$ influences little the $\\nue \\rightarrow \\nus$\\ transitions. In contrast, due to the neutrality and approximate isotopic symmetry of the Earth matter, one has in the Earth: $n_{e}(r) - 0.5~n_{n}(r) \\cong 0.5~n_{e}(r)$. This difference between the number density distributions $n_{e}(r)$ and ($n_{e}(r) - 0.5~n_{n}(r)$) in the Earth is at the origin of the dramatic difference between the magnitudes of the D-N asymmetries corresponding to the small mixing angle $\\nue \\rightarrow \\nus$\\ and $\\nue \\rightarrow \\numt$\\ transition solutions discussed above. Correspondingly, it leads to a shift towards smaller (by a factor of $\\sim 2$) values of \\dms\\ and larger values of \\SdTvS\\ of the iso - \\daynight\\ asymmetry contours in the $\\dms - \\sin^22\\theta_V$ plane corresponding to the $\\nue \\rightarrow \\nus$\\ solution with respect to the analogous contours for the $\\nue \\rightarrow \\numt$\\ solution (compare Figs. 3a - 3c in \\cite{ArticleI} with Figs. 5a, 6a and 7a). At small mixing angles even the \\core\\ asymmetry corresponding to the $\\nue \\rightarrow \\nus$\\ solution is rather small (Table II, Fig. 6a): for $0.0012~\\ltap~ \\sin^22\\theta_V ~\\ltap~ 0.008$ and $2.8\\times 10^{-6}~{\\rm eV^2}~ \\ltap ~\\dms ~\\ltap 4\\times 10^{-6}~{\\rm eV^2}$ we find $(-2\\%)~\\ltap ~\\AsymRC (sterile)~\\ltap ~(-1\\%)$. For other values of \\dms\\ from the small mixing angle ``conservative'' solution region $0.001~ \\ltap~\\sin^22\\theta_V ~\\ltap ~0.009$ one obtains $|\\AsymRC (sterile)| \\leq 1\\%$. We have $\\AsymRC (sterile)~\\gtap ~1\\%$ in the solution region $\\sin^22\\theta_V ~\\gtap~ 0.009$ and $3.0\\times 10^{-6}~{\\rm eV^2}~ \\ltap ~\\dms ~\\ltap ~4.4\\times 10^{-6}~{\\rm eV^2}$. In addition, \\AsymRC (sterile)\\ has a minimum in the interval $0.008~\\ltap~\\sin^22\\theta_V ~\\ltap ~0.03$ at $\\dms \\cong 6.0\\times 10^{-6}~{\\rm eV^2}$ and for this value of \\dms\\ one has $\\AsymRC \\geq 1\\%$ only when $\\sin^22\\theta_V \\geq 0.012$. The \\night\\ and \\mantle\\ asymmetries are larger than 1\\% in absolute value only if $\\sin^22\\theta > 0.010$ (Table II, Figs. 5a and 7a). Replacing the threshold energy $T_{e,th} = 5~$MeV with 7.5 MeV can, depending on the SMA solution value of \\dms, increase $|\\AsymRC|$ (by a factor $\\sim (1.2 - 1.5)$), decrease it somewhat or leave the asymmetry practically the same; it changes little the magnitudes of \\AsymRN\\ and \\AsymRM (Tables III and IV, Figs. 5b, 6c and 7b). The asymmetries \\AsymRC\\ and \\AsymRN, however, are rather sensitive to the value of \\Ye (core) (Tables II - IV and Figs. 5a, 5b and 6a - 6d). The dependence of \\AsymRC\\ and \\AsymRN\\ on \\Ye (core) is particularly strong in the ``conservative'' solution interval $0.0075 ~\\ltap ~\\sin^22\\theta_V ~ < ~0.030$, where a change of the value of \\Ye (core)\\ from 0.467 to 0.50 leads to an increase of $|\\AsymRC|$ and $|\\AsymRN|$ by factors of $\\sim ( 2 - 4)$. The predicted D-N asymmetries in the recoil-e$^{-}$ spectrum for the three samples of events are small in the SMA solution region (Figs. 4.1 - 4.16). The spectrum asymmetry for the \\night\\ sample, for instance, at $\\SdTvS < 0.014$ satisfies $|\\AsymSN(\\Te)|~\\ltap~ 1\\%$ for $5~{\\rm MeV} \\leq T_e \\leq 14~{\\rm MeV}$, and is hardly observable with the Super-Kamiokande detector. This conclusion is valid both for $Y_e(core) = 0.467$ and $Y_e(core) = 0.50$. Analogous results are valid for the \\core\\ sample spectrum asymmetry \\AsymSC (\\Te): one has $|\\AsymSC(\\Te)| \\geq 4\\%$ only if $\\SdTvS \\geq 0.01$; at $\\SdTvS \\cong 0.014$ the asymmetry \\AsymSC(\\Te) reaches 16\\%. The upper limit on the D-N asymmetry \\AsymRN\\ following from the Super-Kamiokande data (eq. (7)) rules out (at 95\\% C.L.) the ``conservative'' large mixing angle (adiabatic) solution possible in the case of solar $\\nue \\rightarrow \\nus$\\ transitions for unrealistically large values of the $^{8}$B neutrino flux \\cite{KPL96}. A qualitative analysis performed by us indicates that the measurement of the \\core\\ and \\mantle\\ sample asymmetries, which are independent observables, can help to discriminate between the $\\nue \\rightarrow \\numt$\\ and the $\\nue \\rightarrow \\nus$\\ transition solutions of the solar neutrino problem. The results obtained in the present study suggest that it will be difficult to probe the small mixing angle nonadiabatic $\\nue \\rightarrow \\nus$\\ transition solution of the solar neutrino problem at $\\SdTvS ~\\ltap~ 0.01$ by performing high precision measurements of the event rate and the recoil-e$^{-}$ spectrum D-N asymmetries with the Super-Kamiokande detector. The precision required to test the indicated solution region exceeds, for most values of the parameters \\dms\\ and $\\SdTvS$ from the region, the precision in the D-N asymmetry measurements which is planned to be achieved in the Super-Kamiokande experiment." }, "9803/astro-ph9803309_arXiv.txt": { "abstract": "We discuss the emergent spectra from accreting black holes, considering in particular the case where the accretion is characterized by relativistic bulk motion. We suggest that such accretion is likely to occur in a wide variety of black hole environments, where the strong gravitational field is expected to dominate the pressure forces, and that this likely to lead to a characteristic high-energy spectroscopic signature; an extended power-law tail. It is in the high (soft) state that matter impinging upon the event horizon can be viewed directly, and the intrinsic power-law is seen. Certain types of Active Galactic Nuclei (AGN) may represent the extragalactic analog of the high-soft state accretion, which would further support our ideas, demonstrating the stability of the ($\\alpha\\sim1.8$) power-law. This stability is due to the asymptotic independence of the spectral index on the mass accretion rate and its weak dependence on plasma temperatures. We have computed the expected spectral energy distribution for an accreting black hole binary in terms of our three model parameters: the disk color temperature, a geometric factor related to the illumination of the black hole site by the disk and a spectral index related to the efficiency of the bulk motion upscattering. We emphasize that this is a fully self-consistent approach, and is not to be confused with the more common phenomenological methods employing additive power law and black-body or multi-color disk. A test of the model is presented using observational data from the Compton Gamma Ray Observatory and the Rossi X-Ray Timing Explorer, covering $\\simeq 2-200$ keV for two recent galactic black hole X--ray nova outbursts. The resulting model fits are encouraging and, along with some observational trends cited from the literature, they support our bulk-motion hypothesis. ", "introduction": "Do black holes interact with an accretion flow in such a way that a unique observational signature can be identified -- that is one which is entirely distinct from those associated with other compact objects, based solely on the radiation observed at infinity? This is a crucial question confronting both theoretical and observational astrophysicists today; for recent reviews of the astrophysics of black holes [see e.g., \\cite{liang97}, \\cite{zhang97a}]. Certainly a large body of evidence has been accumulated which supports the {\\it existence} of black holes, the most convincing arguments being those invoking dynamical mass determinations [e.g. \\cite{orosz97}]. Other arguments have recently been advanced suggesting that X-ray nova flux histories demonstrate the existence of black-hole horizons [\\cite{narayan96}], and in AGN, asymmetrical line features have identified and interpreted as originating in massive black hole environments [\\cite{tanaka95}, \\cite{fabian95}]. Also, quasi-periodic oscillations (QPOs) have been attributed to dynamical time scales associated with the innermost stable orbits in black hole binaries [\\cite{mrg97}]. A distinct feature of black hole spacetime geometries, as opposed to those associated with other compact objects, is the presence of the event horizon. Near the horizon the strong gravitational field is expected to dominate the pressure forces and thus drive the accreting material into free fall. In contrast, for other compact objects the pressure forces are dominant near the surface and the free fall state is absent. Recently, Titarchuk and Zannias (1998) (hereafter TZ98) have developed the relativistic radiative transfer theory demonstrating that high-energy photons are produced by upscattering from the converging inflow within a few Schwarzschild radii. Only some fraction of the radiation emitted by the accretion disk illuminates the converging inflow site. It can be such a situation that the radiation density (or pressure) determined by the injected energy of those soft disk photons and by the weak amplification they experience [\\cite{tmk97}; hereafter TMK97] is much smaller than the Eddington value. {\\it We argue that then this difference is crucial and it results in a unique observational signature for accreting black holes}. \\par \\noindent As explained above, this signature originates from upscattering of low energy photons by fast moving electrons with velocities, $v$, approaching the speed of light, $c$. A soft photon of energy $E$, in the process of multiple scattering off the electrons, gets substantially blue-shifted to energy \\begin{equation} E^{\\prime}=E{{1-(v/c)\\cos\\theta}\\over{1-(v/c)\\cos\\theta^{\\prime}}} \\end{equation} \\noindent due to Doppler effect provided at least one photon is scattered in the direction of electron motion (i.e. when $\\cos\\theta^{\\prime}\\approx 1$). For example, in the first scattering event we assume the direction of incident photon, $\\theta_1$, is nearly normal to the electron velocity, and the direction of the scattered photon is nearly aligned with the electron velocity. In the process the its outward propagation through the converging-inflow medium, the angle between the photon and electron velocity increases. Thus, in the second event the cosine angle, $\\cos\\theta_2$, tends to approach zero. The angle of outgoing photon, $\\theta_2^\\prime$, has to be large enough, in order for the Doppler boosted photon to reach an observer. Any system having a disk structure around a compact object is expected to have a source of low-energy photons. {\\it The boosted photon component is seen as the extended power-law at energies much higher than the characteristic energy of the soft photons. And it is entirely independent of the initial spectral and spatial distributions of the low-energy photons.} The spectral index of the boosted photon distribution is determined only by the mass accretion rate and the plasma temperature of the bulk flow. {\\it The presence of this high-energy power-law component is a generic feature of the model.} A key ingredient in support of our claim comes from the exact relativistic transfer calculations describing the Compton scattering of the low-energy radiation field of the Maxwellian distribution of fast moving electrons (TZ98). It was proven mathematically that the power law is always present as a part of the black hole spectrum over a wide energy range, extending up to 500 keV. A turnover in the spectrum at about this energy, i.e. at E$\\ltorder m_ec^2$, is a prediction of our model. Other extended power-law components, which may be related to the relativistic electron motion, e.g. in a jet, are not uniquely constrained to this energy band because they are not tied with the electron rest mass $m_ec ^2$. The observations with CGRO/OSSE could in principle confirm or refute this prediction. In practice, the data thus far obtained are signal-to-noise limited and cannot address this issue in a definitive manner. \\par \\noindent In this letter we extract the most important points regarding the radiative transfer and present observational evidence which supports the model. \\section { Bulk-Motion Spectral Models} It has been shown elsewhere (TMK97, TZ98, \\cite{ct95}; hereafter CT95) that two effects, the bulk motion upscattering and the Compton (recoil) downscattering (herein BMC), compete forming the hard tail of the spectrum as an extended power law. The soft part of the spectrum comes from the disk photons seen directly and a subset of those photons which escape from the BMC atmosphere after undergoing a few scatterings but without any significant energy change. It has also been shown that without taking into account special and general relativistic effects, one is able to reproduce the main features of the full relativistic formalism: the overall spectral energy distribution and the dependence of the high-energy power on mass-accretion rate (TMK97; TZ98); also, refer to recent calculations by Laurent \\& Titarchuk (1998) (hereafter LT98). In the relativistic treatment, the Compton downscattering becomes less efficient at high energies, due to Klein-Nishina effects. Also, there is a possibility that the electron distribution in the converging inflow can deviate from a Maxwellian, flattening at high velocities, since there is insufficient time for it to thermalize. The hard photon power-law thus extends to higher energies. At the same time however, the spectrum is steepened as a result of gravitational redshift effects. We have assumed that there is an external illumination of the converging flow by the low-energy black body radiation of an accretion disk having a characteristic temperature $T_{c}$. Furthermore we have assumed that this illuminating radiation impinges on the BMC atmosphere with a certain geometry, which we have paramaterized in terms of a ''fraction\" $ f$. This fraction is really the first expansion coefficient of the spatial source photon distribution over the set of the eigenfunctions of the BMC formulation (TMK97, Eq. 30). As we mentioned above (\\S 1) the spectral index is independent of the illumination fraction, $f$. It is clearly demonstrated in TMK97 (Figs 4). We remind the reader that all reasonable theoretical spectra must exhibit a smooth transition from blackbody-like spectrum to a pure power-law, typically, at energies in the 5--12 keV range for stellar black holes. In the soft state when the accretion rate is higher, the soft photons from the the Keplerian disk cool the hot region (Compton cloud) due to thermal Comptonization and free-free emission (CT95). The cooler converging inflow, as it rushes towards the black hole, scatter the soft-photons within the a radius, $r\\sim \\dot m r_s$ -- some of the photons then undergo outward radiative diffusion. Here $\\dot m=\\dot M/\\dot M_E$, $r_s$ is Scwarzschild radius, $\\dot M$ is the net accretion rate (including accretion from the disk plus any halo or other non-keplerian component), $\\dot M_E \\equiv L_E/c^2=4\\pi GMm_p/ \\sigma_Tc~$ is the Eddington accretion rate, $M$ is the mass of the central object, $m_p$ is the proton mass and $G$ is the gravitational constant. It transfers its momentum to the soft-photons to produce the power-law component extending to energies comparable to the kinetic energy of electrons in the converging inflow, i.e. of order $m_ec^2$. On the other hand {\\it in the hard state, the hot emission cloud covering the BMC zone prevents us from seeing the photons that are upscattered to subrelativistic energies within a few Scwarzschild radii}. The luminosity of the upscatterd component, the hard power law, has to be very small compared to the Eddington luminosity in order for the BMC model to be valid. The relative normalization of the soft component to the hard power-law is less important provided the inferred luminosity of the hard power-law remains consistent with the assumption of negligible radiation pressure near the black-hole horizon. The BMC spectral model can be described as the sum of a thermal (disk) component and the convolution of some fraction of this component $g(E_0)$ with the upscattering Green's function $I(E,E_0)$ (TMK97, Eq. 30). The Green's function has the form of a broken power-law with spectral indices $\\alpha$ and $\\alpha+\\zeta$ for high $E\\geq E_0$ and low $E\\leq E_0$ energy parts respectively, \\begin{equation} F_{\\nu} (E)=\\int_0^{\\infty}I(E,E_0)g(E_0)dE_0. \\end{equation} \\par \\noindent The above convolution is insensitive to the value of the Green's function spectral index $\\alpha+\\zeta$, which is always much greater than one. TZ98 presented rigorous proof that the hard power-law tail is a signature of a Schwarzschild black hole. Furthermore, the same statement is valid for the case of a rotating (Kerr) black hole, although a higher mass accretion rate is required to provide the same efficiency for the soft photon upscattering. We note that the processes of absorption and emission (as free-free or synchrotron radiation) can be neglected provided the plasma temperature of the bulk flow is of order 1 keV or greater for characteristic number densities of order $10^{18}$ cm$^{-3}$ and for magnetic field strengths in the proximity of the black hole of order $10^4-10^5$ gauss or less (CT95). \\section {Application to Recent High Energy Observations} As a test of the model we have collected data resulting from high-energy observations covering recent activity periods in two galactic X--ray novae: GRO J1655--40 [\\cite{zhang97b}] and GRS 1915+105 [\\cite{chaty96}]. X--ray novae comprise perhaps the best test case of the methodology described here, since they are in low-mass binary systems -- avoiding the added complications which may arise from the OB star winds in high-mass binary BHCs such as Cygnus X--1 -- and because they become exceptionally bright in outburst exhibiting frequent and pronounced high-energy spectral state transitions [e.g. \\cite{csl97}, \\cite{ebisawa94} \\cite{esin97}]. Furthermore, as a group they comprise the most convincing galactic black--hole candidates. We constructed composite high-energy spectra for GRO J1655--40 during an outburst in the spring of 1996 (\\cite{hynes98}), covering the ~2-200 keV spectral region and fit these data by the BMC model. This was accomplished using summed, standard mode (128 channel) data from the RXTE/PCA and the 16-channel BATSE/LAD earth-occultation data bracketting the pointed observations. In addition, there was substantial outburst activity in GRS 1915+105 during the latter part of 1996 [e.g. \\cite{bandy98}]. We utilized some of the available data from the same instruments for this event as well. The BMC model described in section 2 was imported into the ''XSPEC\" software package which was used to perform all of the model fitting described here. Our resulting fits are shown in Figure 1. For GRO J1655-40 we obtained a blackbody color temperature of $kT_{c}=1.1\\pm0.1$~keV for the soft photon source, a energy spectral index of $\\alpha=1.60\\pm0.03$, and a geometric factor $f$, parameterizing the fraction of the total soft photon flux illuminating the BMC inflow atmosphere, of $f=0.32\\pm0.02$. The observed 2-100~keV flux was $5.7\\times 10^{-8}$ ergs/cm$^2$/s. We note that the corresponding luminosity in the hard power-law component is about $1.5\\%$ of $L_{E}$ (with an assumption of the distance to the source, $3.2$ kpc and the mass of the central object, 7 solar masses), which is consistent with our assumption of negligible radiation pressure near the event horizon. Similar results; $kT_{c}=0.9\\pm0.1$, $\\alpha=1.68\\pm0.03$ and $f=0.72\\pm0.02$, were obtained for GRS 1915+105. From the inferred 2-200~keV luminosity for GRO~J1655-40, $\\sim 5\\%$ of $L_{E}$, we derive a mass accretion rate (in Eddington units) of order 1, bearing in mind the efficiency of gravitational to radiative energy conversion is of order 5\\% or less, e.g. Shakura \\& Sunyaev 1973 (hereafter SS73). This value of $\\dot M$ is consistent with expected values within the BMC framework. This suggests that the line-of-sight column density of the BMC atmosphere is of order $10^{24}$ cm$^{-2}$ (see, TMK97, Eq. 2). However, because the best-fitted color temperature is about 1 keV, (and this is a lower bound on the BMC plasma temperature) we conclude that the detected X-ray spectrum is not significantly modified by absorption (see also \\S 2). The temperatures we infer for these two sources are somewhat lower than values previously reported [e.g. Zhang et al. (1997b)], however this is to be expected. As noted, our procedure represents a fully self-consistent model deconvolution, whereas most previous approaches are phenomenological, i.e. power law plus black body or multicolor disk. Mathematically, one expects the power law component contribute significant soft-energy flux with the net effect of skewing the thermal residual to higher apparent temperatures. This will not occur with approach, as the hard power law turns over towards low energies (see Fig 3, TMK97). The inferred spectral indices also agree extremely well with our model predictions. In TZ98, calculations of the $\\dot m - \\alpha$ relationship were presented. For the low temperature limit, an asymptotic lower limit of $\\alpha\\simeq1.8$ was calculated; for the higher BMC plasma temperatures (of order 10 keV) this limit is significantly lower, $\\sim1$, and for mass accretion rates of $\\dot m\\sim1$ (see below), $\\alpha$ is precisely in the $1.5-1.8$ range we find (LT98, Titarchuk 1998). Thus, we feel our observational test provides extremely encouraging support of our methodology. Using our inferred color temperature $T_c$ and spectral index $\\alpha$, along with the measured flux normalization, we can estimate the mass accretion rate, black hole mass and source distance within the framework of standard accretion disk theory (e.g. SS73). This additionally requires certain assumptions regarding a ''hardening factor\" -- the ratio of color temperature to the effective plasma temperature (\\cite{shimura95}, hereafter ShT95). In fact, the spectral index (TMK97, TZ98, LT98) depends on the mass accretion rate and the plasma temperature; the disk color temperature is $\\propto (\\dot m/m)^{1/4}$ (SS73), and the normalization is $\\propto \\dot m m/d^2$ (where $d$ is the distance to the source). An ideal test case is GRO~J1655-40, since its mass and distance are known to a high degree of accuracy (relative to other BHCs). The distance we infer, $3.8\\pm1.4$ kpc, is consistent (at the $1-\\sigma$ level) with previous determinations (\\cite{Hjellming95}). Also, the black hole mass we calculate can be reconciled with determinations from dynamical studies (\\cite{orosz97}) provided we used the hardening factor 1.9 (ShT95). This assumed a mass accretion rate $\\dot m=3$, which was the value obtained from Monte Carlo simulations performed to calculate the spectral index dependence on the mass accretion rate and plasma temperature (LT98). Again, this is an encouraging result suggesting that with further refinement, one has a method of mass and distance determination independent of the conventional quiescent spectroscopic and photometric studies, which are not always plausible. ", "conclusions": "The successful application of our basic model to observational data and the inferred physical parameters of those systems in comparison to independent determinations is encouraging. We postulate that (i){\\it The soft state detected in GRO J1655-40 and GRS 1915+105 represents a generic feature of accreting Galactic black holes. An extragalactic analog may now be evident in the Narrow Line Seyfert 1 (NLS1) galaxy population.} (ii) {\\it The BMC spectra represent a characteristic signature of black hole horizons. The disk flux, of order 5\\% $L_{edd}$ tends to cool the ambient environment and the generic hard power-law components is seen by the observer.} (iii) {\\it Because this spectral feature is formed very close to the horizon, ($2-3R_s$), the variability timescales of high-energy line and continuum radiation should be associated with the crossing time scale $t_{cross}\\sim10^{-5}M/M_{\\sun}$} s. (iv) {\\it The variability seen in the soft component is not expected to be correlated with the hard component. It is related to the illumination geometry of a small area of the black hole horizon site, whereas the soft radiation seen by the observer directly emanates from a major fraction of the entire disk which comprises a much larger area.} (v) {\\it QPOs emanating from the inner edge of the accretion disk should lead to a pronounced hard- X-ray variability signature, because the seed photons for the converging flow upscattering come from the same inner disk region.} (vi) {\\it The appearance of an additional bump in the energy range 10-20 keV can be explained in terms of downscattering (reflection) effects (ST80) from the inner edge of the accretion disk.} Our conclusions are consistent with various observations of Galactic and extragalactic black hole systems. For example, in NLS1s the X-ray power-law is significantly steeper and its normalization is more variable, with time scale of order $10^4$ s, than in broad-line Seyfert 1 galaxies. This suggests the NLS1s may represent the extragalactic analog of the high-soft state (\\cite{pounds95}, \\cite{brandt97}, \\cite{comastri98}). Several groups have independently reached similar conclusions (CT95, \\cite{pounds95}). We further note that the equivalent widths of Fe features detected tend to be large, in some cases $\\sim500$ eV (\\cite{comastri98}, hereafter C98). It is worth noting that the detection of the strong hydrogen-like iron line is expected if the source of hard energy photons ($>7$ keV) is located inside the the converging inflow region(and alternatively, line radiation could form in the cooler, Compton cloud ambient to the converging inflow region). It is easy to show that the ionization parameter $\\xi=L/(r^2n)\\approx 10^5$ ergs~cm~$s^{-1}$ is a typical value for the converging inflow and it is almost independent of the central object mass. In this case, only the hydrogen-like iron would be expected (Kallman and McCray 1982), which has been confirmed recently by BeppoSAX observations (C98). Another supportive example is the black hole X-ray binary LMC X-3 which appears to always be in the high state. Its hard-tail component varies independently of the soft component [e.g., \\cite{ebisawa93}]. Recent RXTE observations of Cyg X-1 during a state transition [\\cite{cui97}] revealed a striking decrease in the soft-to-hard photon lag times as the source passes from the hard to soft state. This is very strong empirical evidence that the soft seed photons, which comprise a of fraction the disk thermal component, and the hard-X-ray power law emanate from a common compact region, again consistent with our model. The detection of 67 Hz QPOs from GRS~1915+105 by RXTE was recently reported by \\cite{mrg97}. It was clearly demonstrated that this feature is associated with the high energy component visible in the PCA. This can be explained in terms of a QPO in the inner edge of the disk or by $g-$mode disk oscillations occurring within the characteristic radius of 4 $r_s$ [\\cite{tlm98}]. This should lead to variations in the hard spectral component since significant changes in the illumination geometry of the converging inflow site, can occur (TMK97, TZ98). Similar intrepretation can be applied to the 300 Hz QPOs detected by RXTE in GROJ1655--40 [\\cite{rem97}]. In conclusion, we wish to emphasize once again that the observations presented here, along with some observational trends presented by others in the literature, and the relativistic theory prompt us to claim that {\\it we have identified a generic spectral signature black hole accretion.} \\vspace{0.2in} \\centerline{\\bf{ ACKNOWLEDGMENTS}} We wish to acknowledge the anonymous referee for who made a number of useful comments on the initial draft, as well as Jean Swank and Menas Kafatos for discussion and useful suggestions. Portions of this work were supported by the Rossi X--Ray Timing Explorer and Compton Gamma Ray Observatory Guest Observer Programs. L.T. also would like to acknowledge support from NASA grant NAG5-4965. \\newpage" }, "9803/astro-ph9803286_arXiv.txt": { "abstract": "\\rightskip=\\leftskip The Fornax cluster galaxy FCC 35 shows an unusual multiply-peaked integrated \\ion{H}{1} profile (Bureau, Mould \\& Staveley-Smith 1996). We have now observed FCC 35 with the Australia Telescope Compact Array (ATCA) and have found a compact \\ion{H}{1} source with $M_{HI}$ = 2.2 $\\times$ $10^{8}$ $M_{\\odot}$, and a spatially overlapping complex of \\ion{H}{1} gas with the same mass. By combining optical observations with the \\ion{H}{1} data, we are able to identify FCC 35 as a young compact source of star formation with a nearby intergalactic \\ion{H}{1} cloud which is devoid of stars. We classify FCC 35 as a blue compact dwarf (BCD) or \\ion{H}{2} galaxy, having large amounts of neutral hydrogen, very blue colors ($(U-V)$ = 0.1), and a low metallicity spectrum with strong narrow emission lines. Together with the presence of the \\ion{H}{1} cloud, this suggests that FCC 35 is the result of a recent interaction within the Fornax cluster. ", "introduction": "Our interest in FCC 35 began in 1994 when this galaxy was observed as part of an investigation of the Tully-Fisher relation in Fornax (Bureau, Mould, \\& Staveley-Smith 1996; hereafter BMS). Optically, FCC 35 was identified as a member of the Fornax cluster by Ferguson (1989) and classified as a possible BCD/Sm IV. BMS photometry revealed a high surface brightness and an offset nucleus, despite a regular light profile. The 21 cm single-dish observations of FCC 35 showed three distinct \\ion{H}{1} peaks within a 700 km s$^{-1}$ velocity range (see Fig.~\\ref{fig:f1}). There were no other known \\ion{H}{1} sources within the Parkes' beam, so the explanation for the three-peaked profile was unknown. This anomaly served to motivate further studies of FCC 35. Generally, BCDs like FCC 35 are high surface brightness dwarf galaxies which appear to be undergoing an intense period of star formation (Thuan \\& Martin 1981). The cause of these star formation bursts is not well understood, but in many cases interaction is considered a likely mechanism. Interactions can induce rapid star formation (Bushouse 1987) and the resulting internal motions can continue to induce bursts for up to 10$^8$ yrs after the interaction (Noguchi 1991). There are several types of interaction which can trigger the BCD phenomenon. The first involves an encounter between two spiral galaxies and the formation of \\ion{H}{1}-rich tidal tails. BCDs have been observed and modeled to form at the end of these tails which can extend for hundreds of kpc (e.g. Duc et al. 1997; Barnes \\& Hernquist 1992, 1996; Elmegreen, Kaufman \\& Thomasson 1993; Mirabel, Lutz \\& Maza 1991; Schweizer 1978). The surrounding environment is often left in a disordered state for some time after this type of interaction. Another type of starburst-inducing interaction involves an intergalactic \\ion{H}{1} cloud of similar mass to the progenitor galaxy (Taylor, Brinks \\& Skillman 1993, Taylor 1997). Taylor et al. (1995, 1996) completed a survey of relatively isolated BCDs (also called H~{\\footnotesize II} galaxies) and found $\\approx$57\\% of these to have an \\ion{H}{1} companion which could be triggering the star formation burst. Finally, close encounters between two galaxies of different mass often induce star formation in the smaller component, possibly creating a BCD (Ostlin \\& Bergvall 1993; Lacey \\& Silk 1991). Putting all of these possibilities together strongly suggests that the star formation bursts which produce BCDs are indeed due to interactions. In a cluster environment these interactions are much more likely to occur than in the field. The denser the environment, the higher the potential for an interaction among member galaxies. Fornax is a well-studied nearby cluster (d = 18.2 Mpc; Madore et al. 1996) which has one of the highest galaxy volume densities in the Local Supercluster (Held \\& Mould 1994) and more than twice the central surface density of Virgo (Ferguson \\& Sandage 1988). The relatively low velocity dispersion of Fornax ($\\approx$ 400 \\kms) further favors interaction between cluster members. FCC 35 is therefore in an ideal environment to be affected by the mechanisms described above. The ATCA observations of FCC 35, presented here, reveal two distinct \\ion{H}{1} sources; one compact and regular associated with the optical FCC 35, and one extended and irregular with no optical counterpart. The sources overlap spatially but are separated by 140 km s$^{-1}$ in velocity. These observations indicate that FCC 35 has an \\ion{H}{1} companion of comparable mass. This brings us to the various interaction scenarios. The starburst which is now FCC 35 may be due to a previous interaction with this \\ion{H}{1} source or both components could be the result of a spiral-spiral interaction. It is also conceivable that the \\ion{H}{1} companion itself resulted from a gas outflow associated with the star formation burst. Investigating these possibilities is one of the central topics of this paper. In this paper we present a combination of data which helps to reveal the origin and evolution of FCC 35. We discuss the ATCA observations and data reduction in $\\S$2.1 and the optical imaging and spectroscopy in $\\S$2.2. In $\\S$3.0 we present the results of these observations, including the \\ion{H}{1} distribution ($\\S$3.1), \\ion{H}{1} kinematics ($\\S$3.2), stellar distribution ($\\S$3.3), and physical conditions of the gas ($\\S$3.4). Finally, in $\\S$4.0, we discuss these results and their implications for the formation and evolution of dwarf galaxies, in particular with respect to various interaction scenarios. The 21 cm and optical observations together provide a unique source of information on the nature of BCDs and their companions in clusters. ", "conclusions": "\\subsection{The \\ion{H}{1} Cloud} A position-velocity cut through the center of both FCC 35 and the \\ion{H}{1} cloud (Fig.~\\ref{fig:f13}) shows that despite the spatial overlap , the two components are not connected in velocity space. The cloud may either be a unique source within the Fornax cluster or a foreground or background \\ion{H}{1} object. The former seems to be the most likely considering the velocity of the cloud (V$_{r,\\odot}$ = 1658 km s$^{-1}$) and the mean heliocentric velocity of the Fornax cluster ($\\langle{v}\\rangle = 1450 \\pm$ 34 km s$^{-1}$, $\\sigma_v$ = 350 km s$^{-1}$; Held \\& Mould 1994). The Fornax cluster has a central number density of 500 galaxies Mpc$^{-3}$ (Ferguson 1989), and tidal debris from interactions should be expected (e.g. Theuns \\& Warren 1997). Some of the intergalactic material may be primordial, but in a dense cluster it is likely to have arisen from galactic harassment (Moore et al. 1996). This is especially true when considering the proximity of our \\ion{H}{1} cloud to the center of the Fornax cluster ($03^{h}35^{m}, -35.7^{\\circ}$; Ferguson 1989). The interaction which formed the cloud may not have involved the galaxy FCC 35, and the cloud may simply be ``passing by'' at this stage. Its structure could be affected by FCC 35's presence (see Figs.~\\ref{fig:f2} \\&~\\ref{fig:f4}), but this is difficult to confirm due to the irregular kinematics of the cloud and the sensitivity of its inferred structure on the weighting used in the reduction. It is also possible that the \\ion{H}{1} cloud is a remnant of an interaction in which FCC 35 was involved. We note that NGC 1316C is located only 6$^{\\prime}$ away (in projection) from FCC 35 ($\\Delta$V = 150 km s$^{-1}$) and the two could have interacted in the past. It is conceivable that FCC 35 and the cloud were a single object which was ripped apart tidally into two parts of comparable mass. This seems unlikely, however, considering the regular and compact structure of FCC 35. Yet another possibility is that the \\ion{H}{1} cloud {\\em and} FCC 35 formed through an interaction between two spirals. Dwarf galaxies and massive \\ion{H}{1} clouds have been predicted to form (Barnes \\& Hernquist 1992; Elmegreen et al. 1993) and observed forming (Schweizer 1978; Mirabel et al 1991) in the tidal tails which arise from these interactions. The \\ion{H}{1} cloud could then be the remaining gas from a tidal tail. This possibility will be discussed further in the next section. \\subsection{FCC 35} The amount of neutral hydrogen in FCC 35 ($M_{HI}/M_{Tot}$ = 0.5), together with the optical data, presents a picture of a blue compact dwarf (BCD) or \\ion{H}{2} galaxy. Exponential surface brightness profiles are typical of BCDs (see Fig.~\\ref{fig:f10}), as are offset nuclei (Fig.~\\ref{fig:f9}a; Drinkwater \\& Hardy 1991). The spectrum of FCC 35 is also similar to that of the general BCD population, with relatively low metallicity and strong narrow emission lines (Fig.~\\ref{fig:f12}; Izotov et al. 1997; Masegisa et al. 1994; Walsh \\& Roy 1993; Thuan \\& Martin 1981). The strong H${\\alpha}$ and [OIII] emission lines are a signature of the star formation activity in FCC 35, as are the blue colors towards the galaxy's nucleus (see Table~\\ref{tab:t7}). The formation of BCDs is still not understood. Interaction may be responsible for a significant fraction, but there are also explanations based upon an evolutionary sequence among dwarf galaxies. Davies \\& Phillipps (1988) propose a sequence, dI$\\leftrightarrow$BCD$\\leftrightarrow$dE, which involves repeatedly induced star formation bursts and explains the similarities between different types of dwarf galaxies. The trigger of the BCD phenomenon is described by Gordon \\& Gottesman (1981). They find the majority of dwarf irregulars to have an extended \\ion{H}{1} halo and suggest that the infall of this halo fuels the star formation bursts. FCC 35 does have an extended \\ion{H}{1} halo, but the presence of the \\ion{H}{1} cloud suggests that this star formation burst is interaction related. One type of interaction which has been found to produce BCDs and intergalactic \\ion{H}{1} clouds is the interaction between two spiral galaxies. The tidal tails formed as a result of these interactions can extend for hundreds of kpc (Hibbard \\& Van Gorkom 1996) and create objects of up to 10$^9$ M$_{\\odot}$ (Elmegreen et al. 1993). The tidal features often have low mass-to-light ratios (Hibbard \\& Van Gorkom 1996), and the models of Barnes and Hernquist (1992) predict that these objects would have very little dark matter. Elmegreen et al. (1993) also predict that dwarf galaxies formed as a result of this type of interaction should contain old stars from the original disks plus new stars from the interaction-induced star formation bursts. FCC 35's upper limit on the ionized gas abundance (Z $\\leq$ 0.25 Z$_{\\odot}$) is consistent with tidal formation from the outer disk of a spiral galaxy. Indeed, FCC 35 fulfills many of the criteria related to the spiral-spiral interaction scenario (see Table~\\ref{tab:t3}). It has a relatively low $M_{tot}$/$L_V$ and a (corrected) rotation curve (Fig.~\\ref{fig:f8}) which indicates a truncated mass distribution and (presumably) small amounts of dark matter. However, if this is the origin of FCC 35, we would perhaps expect to observe more remnant \\ion{H}{1} in the surrounding region. This was not apparent in the channel maps obtained within the 43$^{\\prime}$ primary beam of the ATCA. FCC 35's star formation burst could also have been induced through an interaction with its closest neighbor, NGC 1316C, which has not been detected in \\ion{H}{1}. However, this scenario appears unlikely when the age of FCC 35's star burst is taken into account. Its color, (U-V)$_{26\\arcsec}$ = 0.1, corresponds to an age of about 10$^{7}$ years (Larson \\& Tinsley 1978). If FCC 35's star formation burst had been directly triggered by an interaction with NGC 1316C, the relative speed of NGC 1316C would need to be at least 3000 km s$^{-1}$. This is excessive given the low velocity dispersion of the Fornax cluster. However, we recall that internal motions resulting from interactions can induce later star formation bursts (Noguchi 1991), so our arguments do not conclusively exclude this possibility or the spiral-spiral interaction scenario. The most plausible interaction-related cause for the star formation burst in FCC 35 is that the \\ion{H}{1} cloud, whatever its origin, has triggered it. Its relative mass and its proximity in space and velocity make it likely that the cloud is interacting with FCC 35. This situation is not uncommon: Taylor et al. (1994) find that galaxies with \\ion{H}{1} companions tend to have a very low mass-to-light ratios and the mass of the companion is often only an order of magnitude smaller than the mass of the galaxy (see also Walter et al. 1997). The relative velocity, projected separation, and masses of the \\ion{H}{1} cloud and FCC 35 clearly show that they are unbound, so it seems likely that they will drift apart as FCC 35 fades into a low surface brightness or irregular dwarf galaxy." }, "9803/astro-ph9803235_arXiv.txt": { "abstract": "We present an unbiased method for evaluating the ranges of ages and metallicities which are allowed by the photometric properties of the stellar populations that dominate the light of early-type galaxies in clusters. The method is based on the analysis of morphologically-classified early-type galaxies in $17$ clusters at redshifts $0.3\\simlt z\\simlt0.9$ and in the nearby Coma cluster using recent stellar population synthesis models that span a wide range of metallicities. We confirm that metallicity effects must play a role in the origin of the slope of the color-magnitude relation for cluster early-type galaxies. We show, however, that the small scatter of the color-magnitude relation out to redshifts $z\\sim1$ does not formally imply a common epoch of major star formation for all early-type galaxies. Instead, it requires that galaxies assembling more recently be on average more metal-rich than older galaxies of similar luminosity. Regardless of the true ages and metallicities of early-type galaxies within the allowed range, their photometric properties and the implied strengths of several commonly used spectral indices are found to be consistent with {\\it apparently} passive evolution of the stellar populations. Also, the implied dependence of the mass-to-light ratio on galaxy luminosity is consistent with the observed trend. The results of our unbiased analysis define the boundaries in age and metallicity that must be satisfied by theoretical studies aimed at explaining the formation and evolution of early-type galaxies in clusters. ", "introduction": "Early-type galaxies in clusters exhibit a linear color-magnitude (CM) relation indicating that bright galaxies are systematically redder than their faint cluster companions (Visvanathan \\& Sandage 1977\\markcite{vs77}). This remarkable relation shows very small scatter ($\\pm 0.05$ magnitude) in high precision photometry of local clusters such as Coma and Virgo (Bower, Lucey \\& Ellis 1992a\\markcite{ble92a}, 1992b, hereafter BLE92) \\markcite{ble92b} and can be extended to clusters at medium-to-high redshift ($0\\simlt z\\simlt 1$) (Ellis et al 1997, Stanford, Eisenhardt \\& Dickinson 1998)\\markcite{el97}\\markcite{sed98}. A first attempt at explaining the universality of the CM relation involves using the age of each galaxy as the main determinant of its color. Ageing stellar populations redden progressively as stars with decreasing initial mass evolve off the main sequence. Therefore, if the colors of cluster galaxies are purely controlled by age, the small scatter about the CM relation implies a nearly synchronous star formation process for all galaxies of a given mass, while the slope of the CM relation implies systematically older ages for more massive galaxies. As shown most recently by Kodama \\& Arimoto (1997\\markcite{ko97}), such a picture is highly unlikely because it does not preserve the slope nor the magnitude range of the CM relation in time. Another important factor that affects the colors of stellar populations is metallicity. At fixed age, a more metal-rich stellar population will appear redder and fainter than a more metal-poor one (e.g., Worthey 1994 \\markcite{wo94}). Hence, increasing metallicity at fixed age has a similar effect on colors as increasing age at fixed metallicity. This is usually referred to as the {\\it age-metallicity degeneracy} (Worthey 1994\\markcite{wo94}). Several studies have shown that CM relation of cluster elliptical galaxies could be primarily driven by metallicity effects (Larson 1974\\markcite{lar74}; Matteuci \\& Tornamb\\'e 1987\\markcite{mator87}; Arimoto \\& Yoshii 1987 \\markcite{ari87}; Bressan, Chiosi \\& Tantalo 1996\\markcite{br96}; Kodama \\& Arimoto 1997\\markcite{ko97}). The physical mechanism usually involved is that of a galactic wind: supernovae-driven winds are expected to be more efficient in ejecting enriched gas, and hence in preventing more metal-rich stars from forming, in low-mass galaxies than in massive galaxies with deeper potential wells. Although age is generally assumed to be the same for all galaxies in these studies, this has not been proven to be an essential requirement. In fact, scenarios in which E/S0 galaxies progressively form by the merging of disk galaxies (Schweizer \\& Seitzer 1992\\markcite{ss92}) in a universe where structure is built via hierarchical clustering also predict that the CM relation is driven primarily by metallicity effects (Kauffmann \\& Charlot 1998\\markcite{kauf98}). Moreover, age effects could be important if, for example, there is sufficiently strong feedback from early galaxy formation to bias the luminous mass distribution of subsequent generations of galaxies by the heating of intergalactic gas. In this paper we present a new, more model-independent approach for evaluating the full range of ages and metallicities allowed by the spectro-photometric properties of early-type galaxies in clusters. The method is based on the construction of age-metallicity diagrams constrained by the colors of early-type galaxies in the nearby Coma cluster and in 17 clusters observed with the {\\it Hubble Space Telescope} ({\\it HST}) at redshifts up to $z\\approx0.9$ (Stanford et al. 1998\\markcite{sed98}). Such an analysis has hitherto been hindered because of the lack of both accurate stellar libraries for different metallicities and reliable morphological information on cluster galaxies at medium-to-high redshifts. Our results can subsequently be reframed into specific theories of galaxy formation, since they will be indispensable for any model that seeks to produce galaxies resembling those actually observed. In \\S2 we present the spectral evolution models used in this paper. The cluster sample is described in \\S3. In \\S4 we construct the age-metallicity diagrams allowed by the observations, and in \\S5 we compute the corresponding ranges in mass-to-light ratio and in several commonly used spectral indices. We discuss our main conclusions in \\S6. ", "conclusions": "We have shown that the tight photometric constraints on early-type galaxies in clusters allow relatively wide ranges of ages and metallicities for the dominant stellar populations. In particular, the small scatter of the CM relation out to redshifts $z\\sim1$ does not necessarily imply a common epoch of star formation for all early-type galaxies. It requires, however, that galaxies assembling more recently be on average more metal-rich than older galaxies of similar luminosity. In this context it is interesting to mention that, based on the spectral indices of nearby E/S0 galaxies, Worthey, Trager \\& Faber (1996\\markcite{wo96}) favor younger ages for more metal-rich galaxies than for metal-poor ones at fixed velocity dispersion. The results of our unbiased analysis therefore define the boundaries in age and metallicity that must be satisfied by theoretical studies aimed at explaining the formation and evolution of early-type galaxies in clusters. The constraints obtained here on the age and metallicity ranges of E/S0 galaxies are consistent with conventional models in which the galaxies all form monolithically in a single giant burst of star formation at high redshift (e.g., Kodama et al. 1998\\markcite{ko98}, and references therein). In fact, this implies that regardless of the true ages and metallicities of early-type galaxies within the allowed range, their photometric properties will always be consistent with {\\it apparently} passive evolution of the stellar populations. As Figure~6 shows, this consistency even extends to spectral index strengths. Our results are also consistent with scenarios in which E/S0 galaxies are formed by the merging of disk galaxies (Schweizer \\& Seitzer 1992\\markcite{ss92}) in a universe where structure is built through hierarchical clustering (Kauffmann 1996\\markcite{kauf96}; Baugh, Cole \\& Frenk 1996\\markcite{ba96}; Kauffmann \\& Charlot 1998\\markcite{kauf98}). For such scenarios, Figure~3 constrains the metallicity and epoch of the last major event of star formation in E/S0 galaxies and their progenitors (see \\S2 and \\S4). The ages and metallicities of cluster E/S0 galaxies predicted by hierarchical models are found to be consistent with these constraints (Kauffmann \\& Charlot 1998\\markcite{kauf98}). To better assess the origin of E/S0 galaxies in clusters one therefore needs to appeal to observational constraints other than their spectro-photometric properties. For example, conventional models of E/S0 galaxy formation are being challenged by the paucity of red galaxies found at high redshifts in deep surveys (Kauffmann, Charlot, \\& White 1997\\markcite{kcw97}; Zepf 1997\\markcite{zepf97}). Also, morphological distinction between E and S0 galaxies and the evolution of the morphology-density relation out to moderate redshifts appear to point to different formation epochs for E and S0 galaxies (Dressler et al. 1997\\markcite{dres97}). The tightness of the CM relation is proof of a stable process in the assembly of cluster early-type galaxies. However, as we move towards greater redshifts, a drastic change is expected at lookback times that approach the formation of the first E/S0 galaxies. This change can arise as a systematic blueing, an increased scatter or a slope flattening in the CM relation (e.g., Arag\\'on-Salamanca et al. 1993; Charlot \\& Silk 1994\\markcite{cs94}; Kauffmann \\& Charlot 1998\\markcite{kauf98}). An interesting question is raised by the presence of morphologically-selected early-type galaxies with very blue colors in clusters at moderate redshifts (\\S3 and \\S4). If these objects are true cluster members, our analysis shows that they could be young metal-poor galaxies that will later join the CM relation. Hence, we need to probe deeper down the galaxy luminosity function in distant clusters in order to assess whether these objects can have any fundamental bearing on the origin of early-type galaxies." }, "9803/astro-ph9803003_arXiv.txt": { "abstract": "We evaluate the effect of screening by bound electron in ${^7}$Be(p,$\\gamma$)$^8$B reaction, where $^7$Be target contains bound electron, in the framework of the adiabatic representation of the three particle problem. A comparison with two other approximations (united atom and folding) is presented. A good agreement between the ``united atom'' approximation and the exact solution is found. We also discuss the screening corrections induced by two K-shell electrons on a $^7$Be target. The bound electron screening effect consequences for $^7$Be and $^8$B solar neutrino fluxes are discussed. ", "introduction": "In recent years, an increasing attention has been devoted to an accurate estimation of electron screening effect for nuclear fusion reactions in stellar plasma and for the interactions of low-energy ion beams with atomic or molecular targets in laboratory experiments (see refs.~\\cite{Gruzinov,Shoppa,Salpeter,Mitler,Carraro,Brown,Langanke,Shaviv} and references therein). In this Letter we present the first quantum mechanical calculation of screening effect by bound electron in \\begin{equation} \\label{creation} {}^7\\mathrm{Be}+\\mathrm{p} \\longrightarrow {}^8\\mathrm{B}+\\gamma \\end{equation} nuclear fusion from the pp-cycle in the Sun. Contribution of this reaction into the the total luminosity of the Sun is negligible small, but it is directly related to the long-standing ``Solar Neutrino Problem'', -- one of the most intriguing issue in the present-day neutrino astrophysics. Standard physics cannot explain an $^{37}$Ar production rate in the Chlorine experiment smaller than that expected from the solar $^8$B neutrino flux measured by both Kamiokande and (with better statistic) Super Kamiokande. GALLEX and SAGE experiments also indicate beryllium neutrino deficit (see the discussion in ref.~\\cite{Innocenti}). One of the most elegant solutions to the solar neutrino anomaly is resonant neutrino flavor conversion in the sun, that is the so-called MSW effect~\\cite{MSW}. It requires an extension of the minimal standard electroweak theory: neutrino masses and neutrino mixing. These neutrino oscillation parameters are determined in a way that can bridge between the predictions of the standard solar models and the solar neutrino observations. Thus, even in the framework of the standard solar model within a hypothesis of neutrino oscillation (and MSW effect), it is apparently needed more precise calculations of nuclear fusion rates in the sun, because they can significantly affect the neutrino oscillation parameters determination. A careful study of electron screening effect on nuclear fusion rates becomes particularly actual in view of expected high accuracy neutrino flux measurements by a number of new large detectors (Super Kamiokande, SNO). The interpretation of forthcoming data requires relevant precise calculations of solar neutrino fluxes and neutrino energy spectrum. Usually, the effects of surrounding plasma on the nuclear fusion are treated in electrostatic screening approximation. This approximation, being classical or quantum, correctly reflects the major properties of a process only for high relative velocity of the colliding nuclei, when electron density in the vicinity of the fusing nuclei remains almost unchanged during the collision. In the case, when relative velocity of the nuclei is much smaller or comparable with the electron one (and this is the case at solar conditions), the electron density changes following any relative configuration of the nuclei, and the electrons have an impact on a kinetic energy shift of the nuclei. It is therefore natural to consider the phenomenon within the framework of the adiabatic approximation, which comes from the well known Born-Oppenheimer (BO) approximation. The BO approximation allows one to treat nuclear motion independently from the electron coordinate, within a new effective potential which depends on the internuclear distance. Since, the considered nuclear velocities are smaller than the boun electron one, corrections to the BO approximation are expected to be negligible. Obviously, the fusing nuclei are from the continuum energy spectrum. An accurate treatment within the adiabatic approximation of the screening effect by electrons from continuum spectrum requires an additional research but the case of bound electrons presents no special problem (see, for example ref.~\\cite{Melezhik}). As it was argued by A.~Dar, G.~Shaviv, and N.~Shaviv~\\cite{Shaviv,Dar}, the commonly accepted Debye-H\\\"uckel theory is not quite adequate for evaluating the screening effect in not-very-dense stars, like the Sun. There is, also, an experimental evidence that this theory does not provide correct answer for the screening~\\cite{Shoppa}. Actually, this fact is of no importance when the screening due to the plasma electrons is by itself rather small. But it is not the case for the low-lying bound electrons which do screen the electric charge of nuclei much effectively, and moreover, the screening effect drastically increases when energy of the fusing ions decreases. The electron screening can have dramatic effects in very dense stellar cores. At low and moderate energies, the fusion cross section of ``bare'' charged nuclei colliding with the relative momenta $p$ in the center-of-mass frame is expressed as (see ref.~\\cite{Lang}): \\begin{equation} \\label{sigma} \\sigma_b(E) = \\frac{S(E)}{E}e^{-2\\pi\\eta}, \\end{equation} where $S(E)$ is the so-called astrophysical factor which incorporates all nuclear features of the process, $E$ is the collision energy of the nuclei, $\\eta = MZ_1Z_2/(m_e a_0 p)$ is the usual Coulomb parameter, $m_e$ and $M$ are the electron and reduced nuclear masses, respectively, and $a_0$ is the hydrogen Bohr radius. The exponential factor originates from the Coulomb wave function of the internuclear motion $\\psi_E^{\\mathrm C}(R)$ at $R = 0$. The screened cross section $\\sigma_s(E)$ differs by the enhancement factor \\begin{equation} \\label{gamma} \\gamma(E)\\equiv \\frac{\\sigma_s}{\\sigma_b} = \\frac{|\\psi_E(0)|^2}{|\\psi_E^{\\mathrm C}(0)|^2}, \\end{equation} where $\\psi_E(R)$ is the wave function of the internuclear motion which accounts for the bound electron. We evaluate the effect of electron screening of $^7$Be nucleus by one bound electron in reaction (\\ref{creation}) in the framework of the adiabatic approximation for three particle problem. This calculation is compared with two relevant approximations, ``united atom'' (UA) and folding approximations which, as we will demonstrate below, give respectively upper and lower estimates for the screening effect. In the framework of the UA approximation we estimate also the screening effect for $^7$Be nucleus with two K-shell electrons. ", "conclusions": "The enhancement factor (\\ref{gamma}) is plotted in fig.~\\ref{enhan} for all three approximations. As it was expected, the UA approximation always overestimate the exact solution, while the folding approximation underestimates it. Nevertheless, it is easy to see that simple UA prescription gives very close values to the exact solution at kinetic energies above $2$ keV. Therefore, the latter can be used not only as a qualitative, but as a good quantitative approximation to the electron screening by bound electrons. The electron screening is dramatic at very low kinetic energies of the nuclei. However, in a plasma, most of the nuclear fusions come at the Gamow peak energy, that is defined by the strong dependence of the nuclear cross section on energy (\\ref{sigma}) and the fast decrease of the exponential particle distribution. This energy is given by: \\[ E_0 = 1.22(Z_1Z_2T_6)^{2/3}(M/M_p)^{1/3}\\;{\\rm keV}, \\] where $T_6 = T/10^6$ K, $M_p$ is the proton mass. In the solar interior at $r_{\\mathrm eff}/R_\\odot = 0.06$, where the $^7$Be and $^8$B neutrino production reaches its maximum~\\cite{BahPin}, the plasma parameters are $T_6 \\approx 14.7$, the electron density $n_e \\approx 7.7/a_{0}^3$, and the Gamow peak energy in reaction (\\ref{creation}) is about $18$ keV. Then $\\gamma(E_0) = 1.1$, that is, there is 10\\% of an enhancement by one bound electron in the bohron production rate. Simple computations within the UA approximation give the screening effect as: \\begin{equation} \\label{screen} \\gamma(E_0) = e^{\\Delta E/kT}. \\end{equation} Then, 10\\% of an enhancement by one bound electron could be easily reproduced just inserting numbers into the formula (\\ref{screen}). One can apply this formula also for a $^7$Be nucleus with two bound electrons. Then, \\[ \\Delta E = \\left(\\chi_1^{\\mathrm{B }}+\\chi_2^{\\mathrm{B }}\\right) - \\left(\\chi_1^{\\mathrm{Be}}+\\chi_2^{\\mathrm{Be}}\\right) = 227.98 \\: \\mathrm{eV}. \\] Here, $\\chi_1^{\\mathrm{B}} = 340.2$~eV and $\\chi_2^{\\mathrm{B}} = 259.4$~eV are, respectively, the fifth and forth ionization potential of the $^8$B atom and $\\chi_1^{\\mathrm{Be}} = 217.72$~eV and $\\chi_2^{\\mathrm{Be}} = 159.9$~eV are, respectively, the forth and third ionization potential of the $^7$Be atom. Thus, two bound electrons enhancement factor is given by $\\gamma(E_0) = 1.196$, i.e. roughly 20\\%. Using the Saha equation Iben, Kalata and Schwartz~\\cite{Iben} calculated the probabilities $f_1$ and $f_2$ that one or two K-shell electrons are associated with any given $^7$Be nucleus. The calculations were perfomed under the assumption of pure Coulomb electron-ion forces, neglecting all excited states and screening. The probabilites found are \\begin{eqnarray*} f_1 &=& \\lambda\\left[1+\\lambda+0.25\\lambda^2 \\exp{\\left(-\\frac{\\Delta_\\chi}{kT}\\right)}\\right]^{-1}, \\\\ f_2 &=& 0.25\\lambda\\exp{\\left(-\\frac{\\Delta_\\chi}{kT}\\right)}f_1, \\end{eqnarray*} where \\[ \\lambda = n_e\\left(\\frac{h^2}{2\\pi m_e kT}\\right)^{3/2} \\exp{\\left(\\frac{\\chi_1^{\\mathrm Be}}{kT}\\right)}. \\] Here $k$ is the Boltzmann's constant, and $\\Delta_{\\chi}=\\chi_1^{\\mathrm{Be}}-\\chi_2^{\\mathrm{Be}}=63.8$~eV. Inserting numbers one can obtain: $f_1 = 30\\%$, and $f_2 = 3\\%$. Using the calculated abundances of $^7$Be ions, one can estimate the thermal averaged screening effect induced by both one and two bound electrons on a $^7$Be nucleus: \\[ \\langle\\gamma\\rangle-1 = 0.30 \\times 0.1 + 0.03 \\times 0.2 \\approx 0.04. \\] In the Standard Solar Model (SSM) the electron capture rate by $^7$Be nucleus is taken to be about 1000 times faster than the proton capture rate~\\cite{BahPin}. Therefore, a small change in $^8$B production rate does not affect significantly the $^7$Be neutrino flux, although it makes a proportional change in $^8$B neutrino flux. Thus, the electron screening by bound electrons has the prompt consequences on bohron neutrino flux. The electron screening by plasma electrons from the continuum spectrum is expected to contribute significantly to the total enhancement factor, since it is proportional to $n_e$, and thus it has to be taken into account in the exact prediction of neutrino flux change. In summary, in the solar interior K-shell bound electrons enhance $^7$Be(p,$\\gamma$)$^8$B rate and increase $^8$B neutrino production rate by of about 4\\%. Therefore, bound electron screening has an effect on the solar $^8$B neutrinos, and acts with the opposite effect to the berrylium neutrinos. The main essence of the electron screening in nuclear fusions is the change in electron density on a nucleus during the collision of the nuclei. This effect can be treated only in a dynamical calculation like the present three particle calculation or the discussed UA approximation. We thank J.~N.~Bahcall, A.~V.~Gruzinov and V.~A.~Naumov for usefull discussions." }, "9803/astro-ph9803145_arXiv.txt": { "abstract": "We report on the results of a multi-wavelength campaign to observe the soft X-ray transient (SXT) and superluminal jet source \\novasco\\ in outburst using \\HST, \\RXTE\\ and \\CGRO\\ together with ground based facilities. This outburst was qualitatively quite different to other SXT outbursts and to previous outbursts of this source. The onset of hard X-ray activity occurred very slowly, over several months and was delayed relative to the soft X-ray rise. During this period, the optical fluxes {\\em declined} steadily. This apparent anti-correlation is not consistent with the standard disc instability model of SXT outbursts, nor is it expected if the optical output is dominated by reprocessed X-rays, as in persistent low mass X-ray binaries. Based on the strength of the 2175\\,\\AA\\ interstellar absorption feature we constrain the reddening to be $\\EBV=1.2\\pm0.1$, a result which is consistent with the known properties of the source and with the strength of interstellar absorption lines. Using this result we find that our dereddened spectra are dominated by a component peaking in the optical with the expected $\\nu^{1/3}$ disc spectrum seen only in the UV. We consider possible interpretations of this spectrum in terms of thermal emission from the outer accretion disc and/or secondary star, both with and without X-ray irradiation, and also as non-thermal optical synchrotron emission from a compact self-absorbed central source. In addition to the prominent \\HeII\\ 4686\\,\\AA\\ line, we see Bowen fluorescence lines of \\NIII\\ and \\OIII, and possible P~Cygni profiles in the UV resonance lines, which can be interpreted in terms of an accretion disc wind. The X-ray spectra broadly resemble the high-soft state commonly seen in black hole candidates, but evolve through two substates. Taken as a whole, the outburst dataset cannot readily be interpreted by any standard model for SXT outbursts. We suggest that many of the characteristics could be interpreted in the context of a model combining X-ray irradiation with the limit cycle disc instability, but with the added ingredient of a very large disc in this long period system. ", "introduction": "Soft X-ray transients (SXTs), also referred to as X-ray novae, \\cite{TS96} are a class of low-mass X-ray binaries (LMXBs) in which long periods of quiescence, typically decades, are punctuated by very dramatic X-ray and optical outbursts, often accompanied by radio activity as well. The most promising models for explaining the outbursts invoke the thermal-viscous limit cycle instability previously developed for cataclysmic variables \\cite{C93}. These have met with some success in explaining the properties of the outbursts \\cite{CCL95} but there remain difficulties (e.g.\\ Lasota, Narayan \\& Yi 1996). Compared to cataclysmic variables, an important effect that must be included in models of SXTs is X-ray irradiation of the disc and/or the secondary star. Irradiation of the disc will change its temperature structure \\cite{TMW90} and may induce delayed reflares (Chen, Livio \\& Gehrels 1993, Mineshige 1994). The SXT \\novasco\\ was discovered in 1994 July when \\GRO\\ Burst and Transient Source Experiment (BATSE) observed it in outburst at a level of 1.1\\,Crab in the 20--200\\,keV energy band \\cite{Ha95}. Since then it has undergone repeated outbursts to a similar level and shown itself to be a very atypical SXT. The outburst history from 1994--5 has been summarised by Tavani et al.\\ \\shortcite{T96}, who draw attention to the contrast between the 1994 outbursts which were radio-loud with apparent superluminal jets observed (Tingay et al.\\ 1995, Hjellming \\& Rupen 1995) and the 1995 outbursts at a similar X-ray flux as in 1994, but radio-quiet. The optical flux from \\novasco\\ is not as well documented, but Orosz, Schaefer \\& Barnes \\shortcite{OSB95} note that optical brightening does not always accompany X-ray outbursts. After a period of apparent quiescence from late 1995 to early 1996, \\novasco\\ went into outburst again in late 1996 April \\cite{R96}. Orosz et al.\\ \\shortcite{O97} observed an optical rise leading the X-ray rise detected by \\XTE\\ by about 6~days. They suggested that this initial behaviour was consistent with the limit-cycle instability. The subsequent X-ray behaviour, however, was not as expected. The soft X-ray flux (2--10\\,keV), as followed by the \\XTE\\ All Sky Monitor (ASM) remained at an approximately constant level for more than 4~months, though with considerable short term variability while the hard X-ray flux (20--200\\,keV) as monitored by \\GRO\\ BATSE was observed to rise very slowly, not reaching its peak until 4~months after the initial dramatic increase in the soft flux. During this period we carried out a series of simultaneous \\HST\\ and \\XTE\\ visits, backed up by ground based observations and \\GRO\\ BATSE data. We present here our spectral analysis. A subsequent paper will address timing issues. First we summarise the current state of knowledge on the properties of \\novasco: the context in which we interpret our results. \\subsection{System parameters} \\label{ParameterSection} The system parameters of \\novasco\\ are the best known of any SXT. They are summarised in Table~\\ref{ParameterTable}. Hjellming \\& Rupen \\shortcite{HR95} estimate the distance from a kinematic model of the jets to be $3.2\\pm0.2$\\,kpc. We also have a lower limit from observations of the 1420\\,MHz interstellar absorption \\cite{T95} of 3.0\\,kpc and an upper limit of 3.5\\,kpc obtained by the method of Mirabel and Rodr\\'{\\i}guez \\shortcite{MR94}. The latter assumes that we can correctly identify the proper motions of the two jets relative to the central source and then only involves the requirement that these proper motions are produced by material moving at no more than the speed of light. These two constraints support the distance estimate of Hjellming \\& Rupen. Other parameters are taken from Orosz \\& Bailyn \\shortcite{OB97} who model the quiescent light curve at a time when the disc is estimated to contribute less than 10 per cent of the V band light. Their deduced mass of $7.02\\pm0.24$\\,M$_{\\sun}$ makes it clear that the compact object in this system is a black hole. We note that an independent parameter determination by van der Hooft et al.\\ \\shortcite{vdH97} yields values consistent with those of Orosz \\& Bailyn \\shortcite{OB97}, although with larger uncertainties. \\begin{table} \\caption{Adopted parameters for \\novasco.} \\label{ParameterTable} \\begin{center} \\begin{tabular}{lc} \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} Distance & $3.2\\pm0.2$\\,kpc \\\\ Period & $2.62157\\pm0.00015$\\,d \\\\ Mass function & $3.24\\pm0.09$ \\\\ Inclination & $69\\fdg50\\pm0\\fdg08$ \\\\ Mass ratio & $2.99\\pm0.08$ \\\\ Primary mass & $7.02\\pm0.22$\\,M$_{\\sun}$ \\\\ \\noalign{\\smallskip} \\hline \\end{tabular} \\end{center} \\end{table} ", "conclusions": "We have obtained a series of co-ordinated optical, UV and X-ray spectra spanning several months of the outburst of an SXT. Although the optical light curve shows the expected decline, it has a spectrum different to that expected on theoretical grounds. Conversely, although the X-ray spectra showed the familiar high state form, the X-ray fluxes continued to rise through the optical decline, contrary to expectations. We have considered various interpretations of the observations and suggest the following possible scenario: The outburst was triggered by a heating wave in the disc, causing an optical rise as the outer disc enters the hot state and a soft X-ray rise subsequently when the material starts to reach the inner disc. Inflow to the inner regions continues to rise for some time as the disc tries to find a steady state. About a month after the initial rise the accretion mode near the black hole changes. This results in a rise in the hard X-ray activity and the X-ray variability as the extended hard power-law component becomes prominent. This change is accompanied by a brief radio flare. While the X-ray activity is still rising, the disc itself is changing in thickness and/or geometry so that irradiation becomes less efficient allowing the cooling wave to move inwards producing the drop in optical flux at a nearly fixed temperature. There is also some irradiation of the secondary star, producing an orbital variation in the continuum fluxes and \\HeII\\ 4686\\,\\AA\\ emission. There clearly remain important unanswered questions not only about \\novasco, but about the outbursts of SXTs in general. There are many theoretical avenues to be explored in seeking an explanation of these observations, especially in the modelling of long period, large disc systems and the exploration of non-thermal models for the optical emission. This work also has many useful lessons for the observer. We have demonstrated the value of co-ordinated, multi-wavelength campaigns in ruling out interpretations which might be suggested by a part of the dataset, but are inconsistent with the whole. We suggest the following priorities for future observations of SXT outbursts: \\begin{enumerate} \\item UV observations are {\\em crucial} to such a campaign for the following reasons. a) It is only in this region that we may be seeing the expected $\\nu^{1/3}$ characteristic accretion disc spectrum in this dataset; identification of this is an important indicator of the disc temperature distribution. b) It is in the UV resonance lines that we see the signature of an accretion disc wind. Higher resolution, higher signal to noise observations (possible only for a less extremely reddened source) will test models of accretion disc winds and allow an estimate of the mass loss rate. c) The 2175\\,\\AA\\ interstellar absorption feature is our best tool in estimating the reddening of these typically highly reddened objects; other measures such as Na~D-lines are not always reliable in these cases. Without a good estimate of the interstellar reddening we cannot determine and hence interpret the intrinsic spectrum. \\item The campaign should include spectra, or at least multi-colour photometry, spanning and adequately sampling several consecutive orbits. This will allow us to separate orbital spectral modulations from random variability and distinguish between emission from an irradiated secondary star and the accretion disc. In the event of the discovery of an unambiguously eclipsing SXT these observations would be of central importance, allowing eclipse mapping of the emission regions. \\item Comprehensive X-ray spectra spanning as wide an energy range as possible should be an integral part of the campaign. We observe the high energy side of a thermal component, but lower energy coverage is required to accurately characterise this component and distinguish between a single temperature blackbody and a multi-colour disc. \\item The campaign should include good red and near infrared coverage to obtain improved characterisation of the long wavelength turnover in the spectrum. This will help in discriminating between thermal and non-thermal emission, which show different long-wavelength limits, and in the thermal case will provide more comprehensive information on the cooler parts of the system. \\end{enumerate}" }, "9803/astro-ph9803102_arXiv.txt": { "abstract": "We present near-infrared (observed frame) spectra of the high-redshift quasar S4{\\ts}0636+68 at $z=3.2$ which was previously thought to be one of a group of ``strong'' \\ion{Fe}{2} emitters (i.e., $F(\\mbox{\\ion{Fe}{2}}{\\ts} \\lambda\\lambda\\mbox{4434--4684})/F({\\rm H}\\beta) > 1$). Our {\\it K}-band spectrum clearly shows emission lines of H$\\beta$ and [\\ion{O}{3}]{\\ts}$\\lambda\\lambda$4959,{\\ts}5007 as well as optical \\ion{Fe}{2} emission. Our computed value of $F(\\mbox{\\ion{Fe}{2} } \\lambda\\lambda\\mbox{4434--4684})/F({\\rm H}\\beta) \\simeq 0.8$ for S4{\\ts}0636+68 is less than previously thought, and in fact is comparable to values found for radio-loud, flat-spectrum, low-$z$ quasars. Therefore S4{\\ts}0636+68 appears not to be a strong optical \\ion{Fe}{2} emitter. Although more than half (5/8) of the high-$z$ quasars observed to date are still classified as strong optical \\ion{Fe}{2} emitters, their \\ion{Fe}{2}/H$\\beta$ ratios, for the most part, follow the same trend as that of low-$z$ quasars, i.e., an anticorrelation in $EW$(\\ion{Fe}{2})/$EW$(H$\\beta$) versus $EW$([\\ion{O}{3}])/$EW$(H$\\beta$), with radio-loud quasars having a mean value of $EW$(\\ion{Fe}{2})/$EW$(H$\\beta$) approximately half that of radio-quiet quasars at comparable values of $EW$([\\ion{O}{3}])/$EW$(H$\\beta$). ", "introduction": "Since optical \\ion{Fe}{2} emission\\footnote{% The \\ion{Fe}{2} emission feature actually extends from the near-UV into the red optical region of the spectrum. However, following previous convention, we use the term ``optical \\ion{Fe}{2} emission'' in this paper to mean the \\ion{Fe}{2} emission near H$\\beta$ (i.e., \\ion{Fe}{2}{\\ts}$\\lambda\\lambda$4434--4684).} is often one of the prominent features in the spectra of Type 1 active galactic nuclei (AGN), it is perhaps not surprising that several observational and theoretical studies have been made to explain the strength of this feature in quasars (e.g. Phillips \\markcite{Phillips77}1977, \\markcite{Phillips78}1978; \\markcite{Kwan81}Kwan \\& Krolik 1981; \\markcite{Netzer83}Netzer \\& Wills 1983; \\markcite{Wills85}Wills {\\it et al.}\\ 1985; \\markcite{Collin-Souffrin88}Collin-Souffrin {\\it et al.}\\ 1988; \\markcite{Zheng90}Zheng \\& O'Brien 1990; \\markcite{Joly91}Joly 1991; \\markcite{Boroson92}Boroson \\& Green 1992; \\markcite{Lipari93}L\\'{\\i}pari {\\it et al.}\\ 1993; \\markcite{Wang96a}Wang {\\it et al.}\\ 1996a). Although it is known that the strength of the optical \\ion{Fe}{2} emission shows an anticorrelation with the strength of [\\ion{O}{3}] emission (Boroson \\& Green 1992), the physical properties of the \\ion{Fe}{2} emitting region are not yet fully understood. Recent near-infrared (NIR) spectroscopic studies by \\markcite{Hill93}Hill {\\it et al.}\\ (1993) and \\markcite{Elston94}Elston {\\it et al.}\\ (1994; hereafter ETH) suggest that unusually strong optical \\ion{Fe}{2} emitters may be common in the high-$z$ universe ($2 < z < 3.4$). Though it is known that some low-$z$ far-infrared (FIR) selected AGN ($L_{\\rm FIR} \\gtrsim 10^{11}$ $L_{\\sun}$) show strong \\ion{Fe}{2} emission in their optical spectra \\markcite{Lipari93}(cf.\\ L\\'{\\i}pari {\\it et al.}\\ 1993), such extreme \\ion{Fe}{2} emitters appear to be rare in the low-$z$ universe. Recently, we obtained NIR spectra of two radio-loud, flat-spectrum, high-$z$ quasars (B 1422+231 at $z=3.6$ and PKS 1937$-$101 at $z=3.8$) and found that their flux ratios of $F(\\mbox{\\ion{Fe}{2} }{\\ts}\\lambda\\lambda\\mbox{4434--4684})/F({\\rm H}\\beta)$ are much less than those of the other high-$z$ quasars (\\markcite{Kawara96}Kawara {\\it et al}.\\ 1996; Taniguchi {\\it et al.}\\ \\markcite{Taniguchi96}1996, \\markcite{Taniguchi97}1997), and in fact are similar to those of radio-loud, flat-spectrum, low-$z$ quasars with normal optical \\ion{Fe}{2} emission. These new observations suggest that high-$z$ quasars may exhibit a range of values of $F(\\mbox{\\ion{Fe}{2} }{\\ts}\\lambda\\lambda\\mbox{4434--4684})/F({\\rm H}\\beta)$ similar to what has been observed for low-$z$ quasars. If the strong \\ion{Fe}{2} emission could be attributed to the overabundance of iron, host galaxies of the high-$z$ quasars with strong \\ion{Fe}{2} emission would form at $z \\sim${\\ts}10 because it is usually considered to be the case that the bulk of the iron arises from Type Ia supernovae which occur $\\sim${\\ts}1--2{\\ts}Gyr after the first major epoch of star formation (e.g., \\markcite{Hamann93}Hamann \\& Ferland 1993, \\markcite{Yoshii96}Yoshii {\\it et al}.\\ 1996). It is therefore important to investigate the chemical properties of high-$z$ quasars systematically. In this paper we present new NIR spectroscopy of S4{\\ts}0636+68, a flat-spectrum radio-loud quasar at $z=3.2$, which is reported in ETH as a very strong iron emitter. Based on our new measurements, we discuss whether the fraction of high-$z$ quasars with strong optical \\ion{Fe}{2} emission is substantially higher than that of low-$z$ quasars. ", "conclusions": "\\subsection{The Rest-Frame UV and Optical Spectra of S4{\\ts}0636+68} Figure \\ref{fig-1} shows the spectra of S4{\\ts}0636+68 (solid line) in the $I\\!H\\!K$ bands together with the Large Bright Quasar Survey (LBQS) composite spectrum (dashed line; \\markcite{Francis91} Francis {\\it et al}.\\ 1991) shifted to $z=3.2$. The atmospheric transmission of Mauna Kea is shown in the upper panel. The spectrum clearly shows H$\\beta$ at 2.04 \\micron{} and a broad ``bump'' of \\ion{Fe}{2} emission between 2.15{\\ts}\\micron{} and 2.23{\\ts}\\micron{}. The spike feature in our spectrum marked by `X' is caused by residual atmospheric absorption. Although ETH did not find evidence for [\\ion{O}{3}]{\\ts}$\\lambda\\lambda$4959,{\\ts}5007 emission lines, our $K$-band spectrum shows their presence (which will be discussed later). No prominent emission lines were found in the $I$-, $J$-, and $H$-band regions of the spectrum. \\ion{Mg}{2}{\\ts}$\\lambda$2798 would be expected to appear at $\\sim$ 1.18 \\micron{} but the $J$-band spectrum (not shown in Figure \\ref{fig-1}) was too noisy to be able to study \\ion{Mg}{2}. Since the efficiency of KSPEC in the $J$-band is not high, we do not use the $J$-band data in this paper. \\placefigure{fig-1} In Figure 2, we compare our results with previous optical-NIR spectroscopic studies of S4{\\ts}0636+68 (\\markcite{Sargent89}Sargent {\\it et al}.\\ 1989; \\markcite{Bechtold94}Bechtold {\\it et al}.\\ 1994; ETH). Their basic data are summarized in Table \\ref{tbl-1}. Our $K$-band spectrum is twice as bright as that of ETH. On the other hand, our $I$-band spectrum is 40{\\ts}\\% fainter than that of \\markcite{Bechtold94}Bechtold {\\it et al}.\\ (1994). These flux discrepancies may be due to possible time variation inherent in the object or to calibration errors in the absolute photometry. However, since the two optical spectra taken at different observing dates over two years are quite consistent with each other (\\markcite{Sargent89}Sargent {\\it et al}.\\ 1989; \\markcite{Bechtold94}Bechtold {\\it et al}.\\ 1994), it seems unlikely that this quasar is highly variable. Therefore, we consider the possibility that the discrepancy may be mostly due to calibration errors. First, we note that our $I\\!H\\!K$ spectra were taken simultaneously and thus there is no internal calibration error in our spectra. Second, the optical spectra of \\markcite{Sargent89}Sargent {\\it et al}.\\ (1989) and \\markcite{Bechtold94}Bechtold {\\it et al}.\\ (1994) show good agreement with each other and thus their photometric calibration seems reliable. Further, the optical power-law slope, $\\alpha=-0.68$ ($f_\\nu \\propto \\nu^\\alpha$), given in \\markcite{Sargent89}Sargent {\\it et al}.\\ (1989) can be consistently extrapolated onto our $K$-band spectrum; the spectral index using the emission-free regions in both the optical spectrum (\\markcite{Sargent89}Sargent {\\it et al}.\\ 1989) and our $K$-band spectrum (1330--1380 \\AA, 1430--1460 \\AA, and 5400--5850 \\AA{} in the rest frame) is estimated to be $\\alpha=-0.69$. Since the seeing during our observations was $\\simeq 0\\farcs{}5$ (FWHM) and our slit size was 1\\arcsec{}, we think that we have detected nearly all of the light from the quasar. Though the details of the observing conditions and slit size are not given in ETH (see Table \\ref{tbl-1}), seeing conditions on Mauna Kea are often better than at KPNO, judging from our experience at KPNO (see \\markcite{Kawara96}Kawara {\\it et al}.\\ 1996; \\markcite{Kawara97}Taniguchi {\\it et al}.\\ 1997), and, therefore we expect that our new measurement is more reliable. \\begin{table} \\dummytable\\label{tbl-1} \\end{table} \\placetable{tbl-1} \\placefigure{fig-2} \\subsection{The Rest-frame Optical Emission-line Properties of S4{\\ts}0636+68} The main aim of our current observations is to provide a more accurate measure of the optical \\ion{Fe}{2}/H$\\beta$ ratio in S4{\\ts}0636+68. Figure \\ref{fig-3} compares our result with that of the ETH. (Note that the flux of ETH spectrum is scaled by a factor of two for proper comparison.) Center positions of H$\\beta$, [\\ion{O}{3}]{\\ts}$\\lambda\\lambda$4959,{\\ts}5007, and the \\ion{Fe}{2} multiplet (42) at a redshift $z=3.2$ are marked in Figure \\ref{fig-3}. The peak positions of H$\\beta$, [\\ion{O}{3}]{\\ts}$\\lambda\\lambda$4959,{\\ts}5007, and \\ion{Fe}{2}{\\ts}$\\lambda$5169 (one of \\ion{Fe}{2} multiplet 42 lines), coincide between the two spectra. The $K$-band spectrum of ETH appears to be dominated by very strong \\ion{Fe}{2} emission with weak, or nondetected [\\ion{O}{3}]{\\ts}$\\lambda\\lambda$4959,{\\ts}5007. ETH actually stated that [\\ion{O}{3}]{\\ts}$\\lambda\\lambda$4959,{\\ts}5007 was not detected, although they noted that their spectrum had small bumps of low significance at the position of the [\\ion{O}{3}] lines. On the other hand, our $K$-band spectrum clearly shows emission peaks which can be identified with [\\ion{O}{3}]{\\ts}$\\lambda\\lambda$4959,{\\ts}5007. \\placefigure{fig-3} To measure the \\ion{Fe}{2} fluxes in our spectrum, we fit emission-line features simultaneously with a least-squares algorithm. Such fitting results depend on the adopted continuum spectrum. As shown in Boroson \\& Green (1992), a local linear continuum is usually adopted to fit H$\\beta$, [\\ion{O}{3}]{\\ts}$\\lambda\\lambda$4959,{\\ts}5007, and the \\ion{Fe}{2} features. However, we have already obtained a global power-law continuum using the rest-frame UV and optical spectra as shown in Figure 2. Therefore, we performed spectral fitting for two cases; 1) local linear continuum and 2) global power-law continuum. In the fitting procedure, we assumed $F(\\mbox{[\\ion{O}{3}]}{\\ts}\\lambda{\\rm 5007})/F(\\mbox{[\\ion{O}{3}]}{\\ts} \\lambda {\\rm 4959})$ = 2.97 (\\markcite{Osterbrock89}Osterbrock 1989). We also assumed that the emission line profiles of H$\\beta$ and the [\\ion{O}{3}]{\\ts}$\\lambda\\lambda$4959,{\\ts}5007 doublet are Gaussian. As for the optical \\ion{Fe}{2} emission features, we used an \\ion{Fe}{2} spectrum of a low-$z$ BAL quasar, PG 0043+039 (\\markcite{Turnshek94}Turnshek {\\it et al}.\\ 1994), as our \\ion{Fe}{2} template. All the emission lines are assumed to have the same redshift. Since it is known that high-ionization broad lines (e.g., \\ion{C}{4}{\\ts}$\\lambda$1549) are often blueshifted with respect to low-ionization lines (Gaskell 1982; Wilkes 1984; Carswell {\\it et al}.\\ 1991; Nishihara {\\it et al}.\\ 1997 and references therein), we use only low ionization lines in our analysis. The fitting results are presented in Figure \\ref{fig-4} and Table \\ref{tbl-2} for each of the two assumed continua. The difference in the line flux ratios between the two cases is less than the measurement errors. Although we do not know which continuum case is more realistic, we adopt the results using the linear continuum fit for further discussion in order to compare our results with those of Boroson \\& Green (1992) and Hill {\\it et al}. (1993) since they also adopted a local linear continuum. In order to examine whether or not the detection of the [\\ion{O}{3}] lines are real in our spectrum, we compare our fit including the [\\ion{O}{3}] doublet with a fit excluding the [\\ion{O}{3}] doublet, where the local linear continuum has been adopted in both fits. A F-statistics test indicates that the fit with [\\ion{O}{3}] is improved over the 4900--5050 \\AA{} region from the fit excluding [\\ion{O}{3}] at a significance level of 99.8{\\ts}\\%. Therefore, we conclude that the ``bumps'' at the [\\ion{O}{3}] positions are really the [\\ion{O}{3}]{\\ts}$\\lambda\\lambda$4959,{\\ts}5007 doublet rather than \\ion{Fe}{2}{\\ts}$\\lambda\\lambda$4924,{\\ts}5018 of the 42 multiplet. We obtained an average redshift $z=3.200 \\pm 0.002$. \\begin{table} \\dummytable\\label{tbl-2} \\end{table} \\placetable{tbl-2} \\placefigure{fig-4} Our fit assuming a local linear continuum gives the flux ratio $F(\\mbox{\\ion{Fe}{2} } \\lambda 5169)/F({\\rm H}\\beta)=0.28$. This value is significantly smaller than the value of 0.45 that we estimate from the published spectrum of ETH. Although we do not understand this difference, it could reasonably be explained by uncertainties in the continuum fits between the two observations. However, we cannot rule out the possibility of time variation. In fact, such a time variation of \\ion{Fe}{2} is reported for the nearby type 1 Seyfert galaxy NGC 5548 (\\markcite{Wamsteker90}Wamsteker {\\it et al}.\\ 1990; \\markcite{Maoz93}Maoz {\\it et al}.\\ 1993; \\markcite{Sergeev97}Sergeev {\\it et al}.\\ 1997). Thus, monitoring of S4{\\ts}0636+68 may be needed in the future. The flux ratio, $F(\\mbox{\\ion{Fe}{2} } \\lambda\\lambda\\mbox{3500--6000})/F({\\rm H}\\beta)$, for S4{\\ts}0636+68 is $3.5\\pm1.1$ (Table \\ref{tbl-2}). This value is greater than the mean value of $1.63 \\pm 0.88$ for the six low-$z$ quasars studied by \\markcite{Wills85}Wills {\\it et al}.\\ (1985) and the value of 2.9 for 3C 273 which is the strongest optical \\ion{Fe}{2} quasar in the sample of \\markcite{Wills85}Wills {\\it et al}.\\ (1985). However, $F(\\mbox{\\ion{Fe}{2}}{\\ts}\\lambda\\lambda\\mbox{4434--4684}) /F({\\rm H}\\beta)$ for S4{\\ts}0636+68 is $0.83 \\pm 0.26$ which is only half of the average value of $1.77 \\pm 0.17$ for the four high-$z$ quasars studied by \\markcite{Hill93}Hill {\\it et al}.\\ (1993). \\markcite{Lipari93}L\\'{\\i}pari {\\it et al}.\\ (1993) defined quasars with $F(\\mbox{\\ion{Fe}{2}}{\\ts}\\lambda\\lambda\\mbox{4434--4684}) /F({\\rm H}\\beta) \\gtrsim 1$ as ``strong'' iron emitters. According to this criterion, we conclude that S4{\\ts}0680+68 is not a strong \\ion{Fe}{2} emitter, contrary to the conclusion of ETH. \\subsection{Statistical Properties of High-$z$ Quasars vs.\\ Low-$z$ Quasars} In order to assess the significance of our new result for the ratio $F(\\mbox{\\ion{Fe}{2}}{\\ts}\\lambda\\lambda\\mbox{4434--4684}) /F({\\rm H}\\beta)$ in S4{\\ts}0680+68 we first compare the rest-frame optical emission line properties of low-$z$ and high-$z$ quasars. Figure \\ref{fig-5} shows the relationship of equivalent width (EW) ratios between $EW(\\mbox{[\\ion{O}{3}]}{\\ts}\\lambda4959+\\lambda5007) / EW({\\rm H}\\beta)$ and $EW(\\mbox{\\ion{Fe}{2} }{\\ts}\\lambda\\lambda\\mbox{4434--4684}) / EW({\\rm H}\\beta)$ for low-$z$ and high-$z$ quasars compiled from the literature (\\markcite{Boroson92}Boroson \\& Green 1992; \\markcite{Hill93}Hill {\\it et al}.\\ 1993; ETH; \\markcite{Kawara96}Kawara {\\it et al}.\\ 1996; \\markcite{Taniguchi97}Taniguchi {\\it et al}.\\ 1997). There is a distinct anticorrelation for the low-$z$ quasars as noted before (cf.\\ \\markcite{Boroson92}Boroson \\& Green 1992) although the reason for the anticorrelation between [\\ion{O}{3}]{\\ts}$\\lambda\\lambda$4959,{\\ts}5007 and optical \\ion{Fe}{2}{\\ts}$\\lambda\\lambda\\mbox{4434--4684}$ is still unknown. The low-$z$ radio-loud quasars tend to have small ratios both in $EW(\\mbox{[\\ion{O}{3}]}{\\ts}\\lambda4959+\\lambda5007) /EW({\\rm H}\\beta)$ and $EW(\\mbox{\\ion{Fe}{2}}{\\ts}\\lambda\\lambda\\mbox{4434--4684}) /EW({\\rm H}\\beta)$. Five of the eight high-$z$ quasars show strong \\ion{Fe}{2} (i.e., \\ion{Fe}{2}/H$\\beta > 1$) emission. The remaining three high-$z$ quasars, which have \\ion{Fe}{2}/H$\\beta < 1$, are all radio-loud and lie within the locus of values traced by low-$z$ radio-loud quasars in Figure 5\\ \\footnote{We note that a radio-loud high-$z$ ($z=2.09$) quasar 1331+170 also appears to lie within the locus of values found for the low-$z$ radio-loud quasars (i.e., 1331+170 appears to have ``quite weak'' optical \\ion{Fe}{2} emission and $EW(\\mbox{[\\ion{O}{3}]}{\\ts}\\lambda4959+\\lambda5007)/EW({\\rm H}\\beta) \\sim{\\ts}0.7$ \\ (Carswell {\\it et al.} \\ 1991).}. Also, the three radio-quiet quasars among the five high-$z$ quasars with strong \\ion{Fe}{2} emission appear to lie within the upper envelope of values observed for low-$z$ radio-quiet quasars. In summary, although over half (5/8) of the high-$z$ quasars appear to be by definition strong \\ion{Fe}{2} emitters, all but two (the radio-loud \\ion{Fe}{2} quasars S5{\\ts}0014+81 and B2{\\ts}1225+317: ETH; \\markcite{Hill93}Hill {\\it et al}.\\ 1993) of the high-$z$ quasars follow a similar trend as that shown by the low-$z$ quasars, i.e.\\ an anticorrelation of $EW(\\mbox{\\ion{Fe}{2}}{\\ts}\\lambda\\lambda\\mbox{4434--4684}) /EW({\\rm H}\\beta)$ versus $EW(\\mbox{[\\ion{O}{3}]}{\\ts}\\lambda4959+\\lambda5007) /EW({\\rm H}\\beta)$, with radio-loud quasars having on average smaller values of $EW(\\mbox{\\ion{Fe}{2}}{\\ts}\\lambda\\lambda\\mbox{4434--4684}) /EW({\\rm H}\\beta)$ than radio-quiet quasars at any given value of $EW(\\mbox{[\\ion{O}{3}]}{\\ts}\\lambda4959+\\lambda5007) /EW({\\rm H}\\beta)$. \\placefigure{fig-5} Recently Wang {\\it et al}.\\ (\\markcite{Wang96b}1996b) studied the relation between optical \\ion{Fe}{2} strength and properties of the UV spectra for 53 low-$z$ ($z \\lesssim 0.2$) quasars and found that there is a significant anticorrelation between the equivalent widths of optical \\ion{Fe}{2}{\\ts}{\\ts}$\\lambda\\lambda\\mbox{4434--4684}$ and \\ion{C}{4}{\\ts}$\\lambda$1549. We examine whether the high-$z$ quasars follow this anticorrelation (Table \\ref{tbl-3} and Figure \\ref{fig-6}). It is perhaps expected that the high-$z$ quasars would have smaller {\\it EW}(\\ion{C}{4}{\\ts}$\\lambda$1549) than low-$z$ quasars simply because of the known anticorrelation between {\\it EW}(\\ion{C}{4}{\\ts}$\\lambda$1549) and UV continuum luminosity (Baldwin effect: \\markcite{Baldwin77}Baldwin 1977; \\markcite{Baldwin78}Baldwin {\\it et al}.\\ 1978). Not as evident perhaps is that, except for 0933+733, the {\\it EW}(\\ion{Fe}{2}{\\ts}$\\lambda\\lambda$4434--4684), appears to show the same range of values as do the low-$z$ quasars at comparable low values of {\\it EW}(\\ion{C}{4}{\\ts}$\\lambda$1549). However, five of the remaining seven high-$z$ quasars (B2{\\ts}1225+317, 1246$-$057, S4{\\ts}0636+68, B{\\ts}1422+231, and PKS{\\ts}1937$-$101) have smaller {\\it EW}(\\ion{Fe}{2}{\\ts}$\\lambda\\lambda$4434--4684) than any of the low-$z$ quasars with comparable {\\it EW}(\\ion{C}{4}{\\ts}$\\lambda$1549) (see the lower-left region of the diagram in Figure 6), thus, adding the high-$z$ sample to the low-$z$ sample appears to decrease somewhat the significance of the anticorrelation between {\\it EW}(\\ion{Fe}{2}{\\ts}$\\lambda\\lambda$4434--4684) versus {\\it EW}(\\ion{C}{4}{\\ts}$\\lambda$1549), (although it is possible that not having a less luminous high-$z$ sample may cause a selection effect). We thus consider it possible that the \\ion{Fe}{2}{\\ts}$\\lambda\\lambda$4434--4684 emitting region may not have a physical link directly with the \\ion{C}{4}{\\ts}$\\lambda$1549 emitting region. However, it seems clear that there is no object with large EWs in both \\ion{Fe}{2}{\\ts}$\\lambda\\lambda$4434--4684 and \\ion{C}{4}{\\ts}$\\lambda$1549, and furthermore, the upper bound of {\\it EW}(\\ion{Fe}{2}{\\ts}$\\lambda\\lambda$4434--4684) still decreases with increasing {\\it EW}(\\ion{C}{4}{\\ts}$\\lambda$1549) for the combined high-$z$ and low-$z$ samples. Hence, it is suggested that there may still be an indirect relation between the \\ion{Fe}{2}{\\ts}$\\lambda\\lambda$4434--4684 and \\ion{C}{4}{\\ts}$\\lambda$1549 regions. \\begin{table} \\dummytable\\label{tbl-3} \\end{table} \\placetable{tbl-3} \\placefigure{fig-6} In summary, the relations among the emission-line properties shown in Figures \\ref{fig-5} and \\ref{fig-6} appear to be valid for both the low-$z$ and high-$z$ quasars with only a few exceptions. This implies that the emission mechanism and the physical properties of the emission-line region in high-$z$ quasars may not be significantly different from those in low-$z$ quasars." }, "9803/astro-ph9803334_arXiv.txt": { "abstract": "We have analyzed the polarization properties of pulsars at an observing frequency of 4.9 GHz. Together with low frequency data, we are able to trace polarization profiles over more than three octaves into an interesting frequency regime. At those high frequencies the polarization properties often undergo important changes such as significant depolarization. A detailed analysis allowed us to identify parameters, which regulate those changes. A significant correlation was found between the integrated degree of polarization and the loss of rotational energy $\\dot E$. The data were also used to review the widely established pulsar profile classification scheme of core- and cone-type beams. We have discovered the existence of pulsars which show a strongly {\\it increasing} degree of circular polarization towards high frequencies. Previously unpublished average polarization profiles, recorded at the 100m Effelsberg radio telescope, are presented for 32 radio pulsars at 4.9 GHz. The data were used to derive polarimetric parameters and emission heights. ", "introduction": "Polarimetry plays a key role in our understanding of the emission mechanism of pulsars, the ambient conditions in the emission region and the geometrical structure of the magnetic field. Similar to the great variety of profile shapes, the polarimetric features of the radio emission vary strongly from pulsar to pulsar and from frequency to frequency. Virtually every polarization state between totally unpolarized and fully linearly or highly circularly polarized can be found amongst different pulsars and even within one profile. Also the shape of the polarization position angle (hereafter PPA) curve varies between nearly constant, a smooth orderly swing, sudden jumps and nearly chaotic behaviour. The jumps in the PPA--swing often cover precisely $90^\\circ$ and are therefore called orthogonal polarization modes (hereafter OPM, see e.g. Stinebring et al. 1984; Gil \\& Lyne 1995; Gangadhara 1997). Nevertheless certain common pulse features have been identified in the past, which lead to different classification attempts (e.g. Backer 1976; Rankin 1983; Lyne \\& Manchester 1988). It is widely accepted that two general types of profile components can be identified: Those which are radiated from the outer parts of the emission tube as {\\it conal} profile components and those which are usually emitted from the central part as {\\it core}-beams. This classification was initially formulated systematically by Rankin (1983), for a detailed description we refer to that paper, a short summary is given in Sect. \\ref{types}. The identification of these components is mainly (but not only) based on the frequency development of their polarization and their relative intensity. In general this system has proven to be remarkably successful, although in this paper we discuss some groups of pulsars which do not quite fit into this classification scheme. One common polarimetric feature of most pulsars is the depolarization towards high frequencies (in the following ``high frequency'' means radio frequencies well above one GHz). Whereas the degree of polarization of pulsars is usually constantly high at low frequencies, it decreases rapidly above a certain frequency (e.g. Manchester 1971; Morris et al. 1981a; Xilouris et al. 1996). This tendency is in contrast to the known properties of other astrophysical objects which have usually stronger polarization at higher frequencies, where the Faraday-depolarization effect is less severe. Therefore, this effect is thought to be inherent to the pulsar magnetosphere, either intrinsic to the emission mechanism or due to a propagation effect within the magnetosphere. The identification of the depolarization mechanism is important as it might help to understand the environmental conditions in the magnetosphere and the relevant physical processes. It is therefore necessary to carry out high frequency observations as we would like to identify parameters which control this effect. In this paper we also focus on the role of the circular polarization. Theories proposed to explain this type of polarization range from purely intrinsic mechanisms (e.g. \\cite{RR90}) to pure propagation effects (e.g. \\cite{MS77}) and combinations of both (e.g. Kazbegi et al. 1991; Naik \\& Kulkarni 1994). In order to distinguish between the different mechanisms, it is necessary to trace the frequency development of the circular polarization over a large frequency interval. Propagational effects should show a strong frequency dependence. Whereas the degree of polarization varies strongly with frequency, the measured PPA is very stable over many octaves in frequency. If one allows for the occurrence of OPMs, the PPA swing is therefore thought to reflect the geometry of the pulsar magnetosphere as first noted by Radhakrishnan \\& Cooke (1969). In some cases it is therefore possible to determine the viewing geometry of a pulsar by fitting the geometry dependent theoretical PPA curve -- the rotating vector model (hereafter RVM) -- to the measured PPA (the formula for the RVM is given e.g. by Manchester \\& Taylor (1977)). Many authors indicate the existence of a radius-to-frequency mapping (hereafter RFM) where the radio emission is narrow band and scales inversely with frequency (e.g. Cordes 1978; Blaskiewicz et al. 1991; Kramer et al. 1997; Kijak \\& Gil 1997; von Hoensbroech \\& Xilouris 1997a). The knowledge of the existence and the strength of the RFM is important as it will help to understand the emission physics. It is therefore necessary to determine the emission height above the pulsar surface, where the emission we observe at a certain frequency, originates. Polarimetry provides one method amongst others to calculate this height. This method was proposed by Blaskiewicz et al. (1991) and we have applied it to our data whenever possible (see Sect. \\ref{Rem}). Due to their steep radio spectra, pulsars tend to be rather weak sources at centimetre wavelengths. As a result, relatively little published data exists in this part of the spectrum (Morris at al. 1981b; Xilouris et al. 1994; Xilouris et al. 1995; Manchester \\& Johnston 1995; Xilouris et al. 1996; von Hoensbroech et al. 1997b). In this paper we present the polarimetric properties of 32 weaker pulsars at 4.9 GHz which roughly doubles the number of published polarization profiles at this frequency and allows statistical studies at such a high frequency for the first time. ", "conclusions": "We have analysed average radio pulsar polarization profiles at high frequencies and present 32 previously unpublished pulsar polarization profiles measured at a frequency of 4.9 GHz. The profiles are also available in EPN-format (\\cite{L98}) through the EPN internet database (see Sect. \\ref{discussion}). The properties of the whole available set of 4.9 GHz average polarization profiles in general and those of individual pulsars in special were compared to lower frequencies. Investigating the average polarization profiles of individual pulsars with particular respect to their classification in the scheme of Rankin (1983), we found groups of pulsars which deserve additional attention. \\begin{itemize} \\item Some pulsars, such as PSR B0355+54, have components with very different polarimetric and spectral properties within the same profile. One component is nearly fully polarized and has a flatter spectrum than the profile as a whole, thus dominates the profile at high frequencies. The rest of the profile is hardly polarized and dominates only at low frequencies. As all eight pulsars which form this group so far, show a similar frequency-dependence, it is suggested that an intrinsic correlation exists between a high degree of polarization and a flat spectrum. All these pulsars have been classified as half-cones. Although this classification is tempting, it is important to note that not a single full-cone with similar properties could be found. This indicates that a more general process takes place than just an occasional lack of flux at the position where the line of sight cuts the cone for the second time. \\item There are three young pulsars (B1800-21, B1823-13 and B1259-63) with a very high loss of rotational energy $\\dot E$ and a very flat spectrum. These pulsars are nearly fully polarized and do not show any significant depolarization. This behaviour shows similarities to the above mentioned highly polarized components of the 0355-like pulsars. Both groups show a correlation between high polarization and a flat spectrum. \\item We found a number of objects which show a circular polarization, which strongly {\\it increases} with frequency. This is in sharp contrast to the known frequency-dependence of pulsar polarization. As the linear polarization decreases simultaneously, it is suggested that a propagation-effect similar to a $\\lambda/4$-plate is active. If confirmed, this could indicate, that propagation effects influence the polarization within the magnetosphere. Additionally, these pulsars fit hardly into the classification scheme. PSR B0144+59 for instance shows precisely the opposite frequency-development to a $S_t$-pulsar (see Fig. \\ref{0144_freq}). \\end{itemize} We would like to point out again the possible role of pulsar evolution on the polarization profile. Within the empirical model for pulsar emission, the profile shape and the polarization properties is nearly exclusively determined by the viewing geometry of pulsar and line of sight and the activity in the different parts of the magnetosphere. But, as it was already noted by Rankin (1983), the classical conal double pulsars (the ``textbook-pulsars'', 0525+21-type) are without exception very old stars which lie close to the ``death-line'' in the $P-\\dot P$-diagram. Contrarily the 1800$-$21-like pulsars mentioned above are very young pulsars with very different properties. For future work it appears to be important to focus stronger on the role of evolution for polarization profile shapes. Analysing the general properties of pulsar polarization profiles at this frequency of 4.9 GHz, we found a significant correlation between the total degree of polarization with $\\dot E$ and the $\\Phi_\\parallel$ at the polar gap respectively (see Fig. \\ref{PEdot}, upper plot). Such a correlation does not exist at lower frequencies. The pulsars at low frequencies rather form groups which have a decreasing (highly polarized, low $\\dot E$ pulsars) and an increasing (weakly polarized, high $\\dot E$ pulsars) degree of polarization to high frequencies. Observations at high radio frequencies therefore yield additional information on the emission physics which is not seen at lower frequencies. This correlation confirms the relation between the depolarization index and $\\Phi_\\parallel$ at 10.5 GHz, which was presented by Xilouris et al. (1995). The large differences in the degree of polarization between individual pulsars indicate that the ambient physical conditions in the respective emitting region differ significantly among them. As we can see from the correlation, this depends on $\\dot E$. $\\dot E$ again is closely correlated to the polar gap $\\Phi_\\parallel$. As a high degree of polarization seems to be correlated to a flatter spectrum (see above), we speculate that a high $\\Phi_\\parallel$ could induce a flatter energy distribution function of the radiating plasma. \\onecolumn \\begin{figure}[C] \\epsfysize23cm \\epsffile[25 70 525 770]{datn1.ps} \\caption{Pulsar polarization profiles at 4.85 GHz. The dark-shaded area represents the linear, the light-shaded area corresponds to the circularly polarized intensity ({\\it positive} $\\hat =$ left-hand, {\\it negative} $\\hat =$ right-hand). Total power is represented by the unshaded solid line. The error-box has a height of 2 $\\sigma$ and a width corresponding to the effective time-resolution (see caption of Table~1). When it was possible, the RVM was fitted to the angle (e.g. for B0144+59).} \\label{data1} \\end{figure} \\twocolumn \\onecolumn \\begin{figure}[C] \\epsfysize23cm \\epsffile[25 70 525 770]{datn2.ps} \\caption{Pulsar polarization profiles at 4.85 GHz. For details see caption of Fig. 8.} \\label{data2} \\end{figure} \\twocolumn \\onecolumn \\begin{figure}[C] \\epsfysize23cm \\epsffile[25 70 525 770]{datn3.ps} \\caption{Pulsar polarization profiles at 4.85 GHz. For details see caption of Fig. 8.} \\label{data3} \\end{figure} \\twocolumn \\onecolumn \\begin{figure}[T] \\epsfysize11.5cm \\epsffile[25 220 525 570]{datn4.ps} \\vspace*{-2cm} \\caption{Pulsar polarization profiles at 4.85 GHz. For details see caption of Fig. 8.} \\label{data4} \\end{figure} \\twocolumn" }, "9803/astro-ph9803044_arXiv.txt": { "abstract": "We have derived the masses of central objects ($M_{\\rm BH}$) of nine type 2 Seyfert nuclei using the observational properties of the {\\it hidden} broad H$\\beta$ emission. We obtain the average dynamical mass, log$(M_{\\rm BH} / M_\\odot) \\simeq 8.00 \\pm 0.51 - 0.475 {\\rm log}(\\tau_{\\rm es}/1)$ where $\\tau_{\\rm es}$ is the optical depth for electron scattering. If $\\tau_{\\rm es} \\sim 1$, this average mass is almost comparable with those of type 1 Seyfert nuclei. However, if $\\tau_{\\rm es} \\ll 1$, as is usually considered, the average mass of type 2 Seyfert nuclei may be more massive than that of type 1s. We discuss implications for issues concerning both the current unified model of Seyfert nuclei and physical conditions of the electron scattering regions. ", "introduction": "It is generally considered that active galactic nuclei (AGNs) are powered by single, accreting supermassive black holes (e.g., Rees 1984; Blandford 1990). According to this scenario, the accretion rate onto a black hole is an important parameter to explain the huge luminosity of AGNs. However, since the luminosity released from this central engine is proportional to the mass of the black hole [ie., the Eddington luminosity, $L_{\\rm Edd} \\sim 10^{46} (M_{\\rm BH}/10^8 M_\\odot)$ erg s$^{-1}$ where $M_{\\rm BH}$ is the black hole mass], the mass itself is considered as another important parameter (Blandford 1990). Relationships between mass and luminosity of AGNs provide important information about the nature of central engines (e.g., Wandel \\& Yahil 1985; Padovani \\& Rafanelli 1988; Padovani 1989; Koratkar \\& Gaskell 1991b). Further, the mass function of nuclei may place constraints on the formation and evolution of supermassive black holes in the universe (Padovani, Burg, \\& Edelson 1990; Haehnelt \\& Rees 1993). Therefore, the mass of AGNs is of fundamental importance in understanding the AGN phenomena. In order to estimate the nuclear mass, the so-called dynamical method has been often used (Dibai 1981, 1984; Wandel $\\&$ Yahil 1985; Wandel \\& Mushotzky 1986; Joly et al. 1985; Reshetnikov 1987; Padovani \\& Rafanelli 1988; Padovani, Burg, \\& Edelson 1990; Koratkar \\& Gaskell 1991b). If the gas motion in a broad emission-line region (BLR) is dominated by the gravitational force exerted by the central massive object, the line width can be used to estimate the mass of the central object given the radial distance of the BLR (Woltier 1959; Setti \\& Woltier 1966). Since the recent elaborate monitoring observations of AGNs have shown that the gas motion in the BLRs is almost dominated by the gravitation (Gaskell 1988; Koratkar \\& Gaskell 1991a, 1991c; Clavel et al. 1991; Peterson 1993; Robinson 1994; Korista et al. 1995; Wanders et al. 1995; Wanders \\& Peterson 1996), the basic assumption in the dynamical method is considered to be robust. In section 2, we discuss the method in detail. All the previous estimates of nuclear mass have been made for type 1 Seyfert nuclei (hereafter S1s) and quasars because the dynamical method needs both the flux and the velocity width of broad line emission. This raises the question ^^ ^^ How massive are type 2 Seyfert nuclei (hereafter S2s) and are they similar to those of S1s ?'' Since the discovery of hidden BLR in the archetypical S2 nucleus of NGC 1068 by Antonucci \\& Miller (1985), it has been considered that S2s are S1s in which the BLR as well as the central engine are hidden from direct view (see, for a review Antonucci 1993). Taking this unified scheme into account, we may expect that there is no systematic difference in the nuclear mass between S1s and S2s. Miller \\& Goodrich (1990; hereafter MG90) made a systematic study of hidden BLRs of high-polarization S2s and found that the properties of the hidden BLRs studied by polarized broad H$\\alpha$ and H$\\beta$ emission are nearly the same as those of S1s in the following respects; equivalent widths, line widths, reddening, and luminosities (see also Tran 1995a). The intrinsic H$\\beta$ luminosities\\footnote{We adopt a Hubble constant $H_0 = 50$ km s$^{-1}$ Mpc$^{-1}$ and a deceleration parameter $q_0 = 0$ throughout this paper.} of the S2s amount to $\\sim 10^{43}$ erg s$^{-1}$. MG90 adopted a typical H$\\beta$ luminosity of $10^{43}$ erg s$^{-1}$ for S1s and thus reached the conclusion that the intrinsic H$\\beta$ luminosities are nearly the same between S1s and S2s. However, the typical H$\\beta$ luminosities of S1s are $\\sim 10^{41}$ - $10^{42}$ erg s$^{-1}$ (Yee 1980; Blumenthal, Keel, \\& Miller 1982; Dahari \\& De Robertis 1988). Therefore, the intrinsic H$\\beta$ luminosities of S2s may be significantly more luminous than those of S1s. However, although this suggests that there is a certain systematic difference between S1s and S2s, it is noted that the intrinsic H$\\beta$ luminosities of S2s depend on the estimates of the optical depth for electron scattering, the degree of {\\it true} polarization, and the covering factor of the scatterers (MG90). Thus the comparison of broad H$\\beta$ luminosities between S1s and S2s must be made carefully. After MG90, several new spectropolarimetric observations of S2s have been published (e.g., Tran, Miller, \\& Kay 1992; Antonucci, Hurt, \\& Miller 1994; Tran 1995a, 1995b). New interpretations on the observed low polarizations have been also presented (Cid Fernandes \\& Terlevich 1995; Heckman et al. 1995; Tran 1995c; Kishimoto 1996, 1997). Therefore, it is interesting to revisit the comparison between the hidden BLRs in S2s and the ordinary BLRs in S1s and then to estimate the dynamical masses of S2 nuclei. ", "conclusions": "We have derived the dynamical masses of S2 nuclei. Comparing them with those of S1s, we have obtained that the nuclear masses of S2s are similar to those of S1s provided that $\\tau_{\\rm es} \\sim 1$ while more massive if $\\tau_{\\rm es} \\ll 1$. For example, if $\\tau_{\\rm es} \\sim 0.1$, the nuclear masses of S2s would be systematically larger by an order of magnitude than those of S1s. We now consider physical conditions in the electron scattering regions, and we discuss some implications for the unified model of Seyfert nuclei. Low optical depths have been adopted in previous works based on spectropolarimetry of S2s; e.g., $\\tau_{\\rm es} \\simeq$ 0.05 - 0.1 (Antonucci \\& Miller 1985; MG90; Miller et al. 1991). As discussed by MG90, if $\\tau_{\\rm es} \\ll 1$, we would observe many AGN with polarizations higher than 50\\% provided that the half opening angle of the ionization cone $\\theta_{\\rm c} = \\theta_{\\rm open}/2 \\sim 30^\\circ$ as is observed for many S2s (Pogge 1989; Wilson \\& Tsvetanov 1994; Schmitt \\& Kinney 1996). However, the highest polarization observed so far is 16\\% (NGC 1068: MG90; Antonucci et al. 1994) and the typical polarization is only several percent for the other S2s (MG90; Tran 1995a). The observed polarization is lower than expected, even after the effect of dilution due to the unpolarized continuum radiation is taken into account (Cid Fernandes \\& Terlevich 1995; Heckman et al 1995; Tran 1995c; Tran, Cohen \\& Goodrich 1995; see also Kishimoto 1996). Therefore, the optical thick condition cannot be ruled out entirely at present. We now discuss the physical characteristics of the electron scattering region. The optical depth for electron scattering is estimated by \\begin{equation} \\tau_{\\rm es} \\sim \\sigma_{\\rm T} {\\overline{n}_{\\rm e}} l_{\\rm eff} \\sim \\sigma_{\\rm T} N_{\\rm e} \\end{equation} where $\\sigma_{\\rm T}$ is the Thomson cross section, 0.66$\\times 10^{-24}$ cm$^2$, ${\\overline{n}_{\\rm e}}$ is the average electron density in the scattering region, $l_{\\rm eff}$ is the effective path length, and $N_{\\rm e}$ is the electron column density. We obtain $N_{\\rm e} \\sim 10^{24}$ cm$^{-2}$ for $\\tau_{\\rm es} \\sim 1$ while $N_{\\rm e} \\sim 10^{23}$ cm$^{-2}$ for $\\tau_{\\rm es} \\sim 0.1$. If the kinetic temperature of free electrons is as high as $\\sim 10^6$ K, significant line broadening would occur due to the scattering (Antonucci \\& Miller 1985; MG90). Since $v \\sim 2000 (T_{\\rm e}/10^5 {\\rm K})$ km s$^{-1}$, it seems reasonable to assume $T_{\\rm e} \\sim 10^5$ K at most (see also Miller et al. 1991). If the gas in the electron scattering region is in pressure equilibrium with the BLR gas ($n_{\\rm e} \\sim 10^9$ cm$^{-3}$ and $T_{\\rm e} \\sim 10^4$ K; Osterbrock 1989), then the average electron density in the scattering region is ${\\overline{n}_{\\rm e}} \\sim 10^8$ cm$^{-3}$, and the effective path lengths are $\\sim 10^{16}$ cm and $\\sim 10^{15}$ cm for $\\tau_{\\rm es} \\sim 1$ and $\\sim 0.1$, respectively. Since we observe the BLR through the scatterers, the scattering regions are located outside the BLRs and thus the radial distance of scatterers is $r_{\\rm e} > 10^{16}$ cm. In the case of NGC 1068, it is observed that the scattering regions are extended to $\\sim$ 100 pc from the nucleus (Capetti et al. 1995a, 1995b). This large value was indeed suspected from the photoionization consideration by Miller et al. (1991). It is therefore considered that the electron scattering regions are located at $r_{\\rm e} \\sim 10^{16}$ - $10^{20}$ cm. The effective radius of scatterers may be different from object to object because it depends also on how we observe the dusty tori in AGNs (i.e., viewing angle dependent; cf. MG90, Miller et al. 1991; Kishimoto 1996, 1997). It is worth noting that the above physical conditions are quite similar to those of warm absorbers probed by X-ray spectroscopy of type 1 AGNs (Halpern 1984; Netzer 1993; Nandra \\& Pounds 1992, 1994; Ptak et al. 1994; Reynolds \\& Fabian 1995; Reynolds et al. 1995; Otani et al. 1996; Reynolds 1997). However, the electron column density inferred in this study, $N_{\\rm e} \\sim 10^{23}$ - $10^{24}$ cm$^{-2}$, is higher by one to two orders of magnitude than those of warm absorbers, $\\sim 10^{22}$ cm$^{-2}$. This seems to be inconsistent with the strict unified model of Seyfert nuclei (Antonucci \\& Miller 1985). We may, however, consider the higher column densities in the S2s as due to the effect of multiple scattering if $\\tau_{\\rm es} \\sim 1$. Another interpretation may be that S2s are gas-richer systematically than S1s (cf. Heckman et al. 1989; Taniguchi 1997). \\vspace{0.5cm} We would like to thank Makoto Kishimoto, Youichi Ohyama, and Takashi Murayama for useful discussion. We also thank the anonymous referee for his/her many useful comments and suggestions which improved this paper significantly. This work was financially supported in part by Grant-in-Aids for the Scientific Research (No. 07044054) of the Japanese Ministry of Education, Culture, Sports, and Science. \\newpage \\begin{table} \\caption{Comparison of the FWHM(H$\\beta$)\\tablenotemark{a} ~ between MG90 and Tran(1995a)} \\vspace{5mm} \\begin{tabular}{ccc} \\tableline \\tableline Galaxy & MG90 & Tran (1995a) \\\\ \\tableline Mrk 3 & 5400 & 6000$\\pm$500 \\\\ Mrk 463E & 3000 & 2770$\\pm$180 \\\\ NGC 7674 & 1500 & 2830$\\pm$150 \\\\ \\tableline \\tablenotetext{a}{In units of km s$^{-1}$.} \\end{tabular} \\end{table} \\begin{table} \\caption{The dynamical masses of Seyfert 2 nuclei} \\vspace{5mm} \\begin{tabular}{ccccccc} \\tableline \\tableline Object & FWHM(H$\\beta_{\\rm b}$) & $L({\\rm H}\\beta_{\\rm b})_{\\rm p}$\\tablenotemark{a} & $P$\\tablenotemark{b} & $M_{\\rm BH} (\\tau_{\\rm es}=1)$ & $M_{\\rm BH} (\\tau_{\\rm es}=0.1)$ & Ref.\\tablenotemark{c} \\\\ & (km s$^{-1}$) & (erg s$^{-1}$) & (\\%) & ($M_{\\odot}$) & ($M_{\\odot}$) & \\\\ \\tableline NGC 1068 & 3030 & $1.11\\times10^{39}$ & 16 & $1.73\\times10^7$ & $5.16\\times10^7$ & 1, 2 \\\\ NGC 7212 & 5420 & $3.40\\times10^{39}$ & 22 & $8.09\\times10^7$ & $2.42\\times10^8$ & 2 \\\\ NGC 7674 & 2830 & $5.36\\times10^{39}$ & 8 & $4.43\\times10^7$ & $1.32\\times10^8$ & 2 \\\\ Mrk 3 & 6000 & $1.00\\times10^{40}$ & 20 & $1.73\\times10^8$ & $5.17\\times10^8$ & 2 \\\\ Mrk 348 & 9350 & $2.95\\times10^{39}$ & 35\\tablenotemark{d} & $1.81\\times10^8$ & $5.39\\times10^8$ & 2 \\\\ Mrk 463E & 2770 & $3.99\\times10^{40}$ & 10 & $9.89\\times10^7$ & $2.95\\times10^8$ & 2 \\\\ Mrk 477 & 4130 & $7.85\\times10^{40}$ & 2 & $6.66\\times10^8$ & $1.96\\times10^9$ & 2 \\\\ Mrk 1210 & 3080 & $2.22\\times10^{39}$ & 15 & $2.56\\times10^7$ & $7.63\\times10^8$ & 2 \\\\ Was 49 & 5860 & $3.30\\times10^{40}$ & 20 & $2.91\\times10^8$ & $8.69\\times10^8$ & 2 \\\\ \\tableline \\tablenotetext{a}{Luminosity of polarized, broad H$\\beta$ emission.} \\tablenotetext{b}{Intrinsic polarization corrected for the unpolarized continuum emission as well as interstellar polarization taken from Tran (1995c).} \\tablenotetext{c}{1. Miller \\& Goodrich 1990; 2. Tran 1995a, 1995c.} \\tablenotetext{d}{The intrinsic polarization for H$\\alpha$ emission.} \\end{tabular} \\end{table} \\begin{table} \\caption{Comparison of the average dynamical masses between S2s and S1s} \\vspace{5mm} \\begin{tabular}{ccc} \\tableline \\tableline Sample & Number & $M_{\\rm BH}$ \\\\ & & ($M_{\\odot}$) \\\\ \\tableline S2 ($\\tau_{\\rm es}=1$) & 9 & 8.00$\\pm$0.51 \\\\ S2 ($\\tau_{\\rm es}=0.1$) & 9 & 8.47$\\pm$0.51 \\\\ S1 (PR88) & 30 & 7.89$\\pm$0.57 \\\\ S1 (PBE90) & 25 & 7.48$\\pm$0.63 \\\\ \\tableline \\end{tabular} \\end{table} \\newpage" }, "9803/astro-ph9803272_arXiv.txt": { "abstract": "To extract reliable cosmic parameters from cosmic microwave background datasets, it is essential to show that the data are not contaminated by residual non-cosmological signals. We describe general statistical approaches to this problem, with an emphasis on the case in which there are two datasets that can be checked for consistency. A first visual step is the Wiener filter mapping from one set of data onto the pixel basis of another. For more quantitative analyses we develop and apply both Bayesian and frequentist techniques. We define the ``contamination parameter'' and advocate the calculation of its probability distribution as a means of examining the consistency of two datasets. The closely related ``probability enhancement factor'' is shown to be a useful statistic for comparison; it is significantly better than a number of $\\chi^2$ quantities we consider. Our methods can be used: internally (between different subsets of a dataset) or externally (between different experiments); for observing regions that completely overlap, partially overlap or overlap not at all; and for observing strategies that differ greatly. We apply the methods to check the consistency (internal and external) of the MSAM92, MSAM94 and Saskatoon Ring datasets. From comparing the two MSAM datasets, we find that the most probable level of contamination is 12\\%, with no contamination only 1.05 times less probable, 50\\% contamination about 8 times less probable and 100\\% contamination strongly ruled out at over $2\\times 10^5$ times less probable. From comparing the 1992 MSAM flight with the Saskatoon data we find the most probable level of contamination to be 50\\%, with no contamination only 1.6 times less probable and 100\\% contamination 13 times less probable. Our methods can also be used to calibrate one experiment off of another. To achieve the best agreement between the Saskatoon and MSAM data we find that the MSAM data should be multiplied by (or Saskatoon data divided by): $1.06^{+0.22}_{-0.26}$. ", "introduction": "The cosmic microwave background (CMB) is black body radiation with a mean temperature of $2.728 \\pm 0.002$ K \\cite{firas}. This mean is modulated by a dipole due to our peculiar motion with respect to the radiation field. If one removes the dipole, the temperature is uniform in every direction to $\\pm 100 \\muK$. Precision measurement of these tiny deviations from isotropy can tell us much about the Universe \\cite{forecast}. Unfortunately, precision measurement of $~100\\muK$ fluctuations is not an easy task. Even given sufficient detector sensitivity and observing time, one still has to contend with many possible contaminants such as side lobe pickup of the $300^\\circ$ Kelvin Earth and atmospheric noise (even from high-altitude balloons). In addition there can be contamination of CMB observations by astrophysical foregrounds. Despite these difficulties there is good reason to believe that, at least for some experiments, the signals observed from sub-orbital platforms are not dominated by contaminants. One of the best reasons for believing this comes from the comparisons that have been done---between FIRS and DMR \\cite{Ganga}, Tenerife and DMR \\cite{Lineweaver}, MSAM and Saskatoon \\cite{nett95}, two years of Python data \\cite{Ruhl}, and two flights of MSAM \\cite{Inman}. Especially for the case when data being compared are from two different instruments, almost the only thing their acquisitions have in common is that they were observing the same piece of sky--each dataset has entirely different sources of systematic error. In addition to confirming the astrophysical origin of the estimated signal, comparison can greatly improve the ability to detect foreground contamination. Perhaps the best evidence for the thermal nature of anisotropy comes from the comparison between the MSAM92 and Saskatoon datasets. Together, these observations span a frequency range from 36 GHz to greater than 170 GHz. In \\cite{nett95} it was found that the spectral index $\\beta$ ($\\delta T \\propto (\\nu/\\nu_0)^\\beta$) is constrained to be $\\beta = -0.1 \\pm 0.2$. For CMB, free-free and dust over this frequency range we expect $\\beta = 0$, $-1.45$ and $2.25$, respectively. The authors conclude that the signals (in the region of overlap) are not dominated by contamination from known astrophysical foregrounds, but are, rather, primarily CMB. We should not let this apparent success fool us into thinking that going to the next level of precision will be easy. There is a big difference in the level of toleration of contaminants when the goal switches from detection to precision measurement. It is likely that there will be significant levels of contamination (from the atmosphere, side lobes, and foregrounds) in future sub-orbital missions. It may be difficult to convincingly demonstrate that contamination is low without comparison. Given the importance of comparison, we feel it is worth improving upon the methods used previously. Past treatments have had to ignore much relevant data, and make uncontrolled approximations. This is due to the fact that generally the two datasets being compared were obtained from instruments observing the sky in different ways. The beam patterns and differencing schemes may differ as in the case of the MSAM/Saskatoon comparison. In \\cite{nett95} one of the MSAM differencing schemes was approximately recreated in software in order to do the comparison. However, no use of software could change the fact that the MSAM and Saskatoon beam patterns, although they have fairly similar full-widths at half-maximum, differ significantly in shape. Even when the differencing schemes and beam patterns are the same, there can still be barriers to a direct comparison. The two MSAM flights took data with essentially the same beam pattern and applied the same differencing, but in this case the direct comparison is frustrated by the fact that the pixels do not all line up exactly. Therefore in \\cite{Inman}, pixels within half a beam width of each other were approximated as being at the same point, and those pixels with no partner from the other dataset within this distance were ignored. Half of the data were lost this way. Here we develop methods of comparing datasets that do not require any information to be thrown away. Differences in demodulation schemes, and effects due to non-overlapping pixels are automatically taken into account. The inevitable price we pay for this is model-dependence. However, we generally expect the model-dependence to be small and indeed find it to be so in the case studies shown here. An extremely useful tool for visual comparison is the Wiener filter. Roughly speaking, it allows us to interpolate the results from one experiment onto the expected results for another experiment that has observed the sky differently. After some notational preliminaries in section II, in section III we introduce the Wiener filter in the context of the probability distribution of the signal, given the data. Also in this section we describe the datasets and apply the Wiener filter to them. When comparing datasets we are testing the consistency of our model of the datasets. We emphasize that meaningful model consistency testing demands the existence of other models with which to compare. Therefore we extend our model of the data to include a possible contaminant and calculate the probability distribution of its amplitude, given the data. We find a more limited extension of the model space to also be useful, in which we only consider one alternative to no contamination: complete contamination. We define the ``probability enhancement factor'' as the logarithm of the ratio of the probability of no contamination to the probability of complete contamination. This Bayesian approach to comparison is described and applied in section IV. In section V we discuss and apply frequentist techniques such as $\\chi^2$ tests. The probability enhancement factor can also be used as the basis for a frequentist test---and it is in fact the well-known likelihood ratio test. We demonstrate that the probability enhancement factor has more discriminatory power than any of the other tests considered. After a further look at the data with the probability enhancement factor in section VI, we discuss the fixing of relative calibration in section VII and possible contamination due to dust in section VII. Finally we summarize our results in section IX. ", "conclusions": "We have demonstrated the usefulness of the Wiener filter for making visual comparisons of datasets. We have emphasized that meaningful consistency testing requires alternative models with which to compare. Thus we have explicitly extended our model of the data to include a possible contaminant and calculated the probability distribution of the amplitude of this contaminant. For purposes of extracting just one number from the comparison we advocate calculating the ratio of the probability of no contamination to the probability of infinite contamination. Viewed as a statistic, we have shown this ``probability enhancement factor'' to be better than various $\\chi^2$ statistics at discriminating between competing hypotheses. The utility of our comparison statistics was shown by exercising them on the MSAM92, MSAM94 and SK95 data. We have found from comparing MSAM92 and MSAM94 that the most probable level of contamination is 12\\%, with zero contamination only 1.05 times less probable, and total contamination over $2\\times 10^5$ times less probable. From comparing MSAM92 and SK95 we have found that the most probable level of contamination is 50\\%, with zero contamination only 1.6 times less probable, and total contamination 13 times less probable. Looking at subsets of the data we find a region at large RA where the SK and MSAM measurements disagree. From IRAS and from the MSAM dust measurements we know that this region is also the dustiest region of the overlap between SK and MSAM. The origin of the discrepancy is unclear and may be due to instrumental artifacts in SK, or foreground contamination of either the SK or MSAM measurements. A revolution is underway in the quality and quantity of CMB data---a revolution generated by the satellites MAP and Planck \\cite{satellites} as well as by a number of balloon and ground-based programs. The amount of data may soon be too large for the type of complete statistical analysis described here. However, any approximate methods developed for extracting the power spectrum or parameters will also be applicable to the statistical procedures introduced here." }, "9803/astro-ph9803208_arXiv.txt": { "abstract": "We present analysis of the shape and radial mass distribution of the E4 galaxy NGC 3923 using archival X-ray data from the {\\sl ROSAT} PSPC and HRI. The X-ray isophotes are significantly elongated with ellipticity $\\epsilon_x=0.15 (0.09-0.21)$ (90\\% confidence) for semi-major axis $a\\sim 10h^{-1}_{70}$ kpc and have position angles aligned with the optical isophotes within the estimated uncertainties. Applying the Geometric Test for dark matter, which is independent of the gas temperature profile, we find that the ellipticities of the PSPC isophotes exceed those predicted if $M\\propto L$ at a marginal significance level of $85\\% (80\\%)$ for oblate (prolate) symmetry. Detailed hydrostatic models of an isothermal gas yield ellipticities for the gravitating matter, $\\epsilon_{mass}=0.35-0.66$ (90\\% confidence), which exceed the intensity weighted ellipticity of the $R$-band optical light, $\\langle \\epsilon_R\\rangle = 0.30$ ($\\epsilon_R^{max}=0.39$). We conclude that mass density profiles with $\\rho\\sim r^{-2}$ are favored over steeper profiles if the gas is essentially isothermal (which is suggested by the PSPC spectrum) and the surface brightness in the central regions $(r\\la 15\\arcsec)$ is not modified substantially by a multi-phase cooling flow, magnetic fields, or discrete sources. We argue that these effects are unlikely to be important for NGC 3923. (The derived $\\epsilon_{mass}$ range is very insensitive to these issues.) Our spatial analysis also indicates that the allowed contribution to the {\\it ROSAT} emission from a population of discrete sources with $\\Sigma_x\\propto\\Sigma_R$ is significantly less than that indicated by the hard spectral component measured by {\\sl ASCA}. ", "introduction": "\\label{intro} The structure of the dark matter halos of galaxies provides important clues to their formation and dynamical evolution (e.g. Sackett 1996; de Zeeuw 1996, 1997). For example, in the Cold Dark Matter (CDM) scenario (e.g. Ostriker 1993) there is evidence that the density profiles of halos have a universal form essentially independent of the halo mass or $\\Omega_0$ (Navarro, Frenk, \\& White 1997; though see Moore et al. 1997). The intrinsic shapes of CDM halos are oblate-triaxial with ellipticities similar to the optical isophotes of elliptical galaxies (e.g. Dubinski 1994). The global shape of a halo also has implications for the mass of a central black hole (e.g. Merritt \\& Quinlan 1997). At present accurate constraints on the intrinsic shapes and density profiles of early-type galaxies are not widely available (e.g. Sackett 1996; Olling \\& Merrifield 1997)\\footnote{The distribution of dark matter in spiral galaxies is also far from being a solved problem -- see, e.g. Broeils \\shortcite{broeils}.}. Stellar dynamical analyses that have incorporated the information contained in high order moments of stellar velocity profiles have made important progress in limiting the uncertainty in the radial distribution of gravitating mass arising from velocity dispersion anisotropy (Rix et al. 1997; Gerhard et al. 1997). However, as indicated by the paucity of such stellar dynamical measurements, the required observations to obtain precise constraints at radii larger than $\\sim R_e$ are extensive, and the modeling techniques to recover the phase-space distribution function are complex. It is also unclear whether this method can provide interesting constraints on the intrinsic shapes since only weak limits on the range of possible shapes have been obtained from analysis of velocity profiles out to $\\sim 2$ $R_e$ (e.g. Statler 1994). Interesting measurements of the ellipticity of the gravitating mass have been obtained for two Polar Ring galaxies (Sackett et al. 1994; Sackett \\& Pogge 1995) and from statistical averaging of known gravitational lenses (e.g. Keeton, Kochanek, \\& Falco 1997), but owing to the rarity of these objects it is possible that the structures of their halos are not representative of most early-type galaxies. Moreover, gravitational lenses, which are biased towards the most massive galaxies, only give relatively crude constraints on the ellipticity and radial mass distribution for any individual system and only on scales similar to the Einstein radius (e.g. Kochanek 1991). The X-ray emission from hot gas in isolated early-type galaxies (Forman, Jones, \\& Tucker 1985; Trinchieri, Fabbiano, \\& Canizares 1986; for a review see Sarazin 1997) probably affords the best means for measuring the shapes and radial mass distributions in these systems (for a review see Buote \\& Canizares 1997b; also see Schechter 1987 and the original application to the analogous problem of the shapes of galaxy clusters by Binney \\& Strimple 1978). The isotropic pressure tensor of the hot gas in early-type galaxies greatly simplifies measurement of the mass distribution over stellar dynamical methods. Moreover, since the shape of the volume X-ray emission traces the shape of the gravitational potential independent of the (typically uncertain) gas temperature profile (Buote \\& Canizares 1994, 1996a), the shape of the mass distribution can be accurately measured in a way that is quite robust to the possible complicating effects of multi-phase cooling flows and magnetic fields (see Buote \\& Canizares 1997b). Presently, X-ray measurements of the mass distributions in early-type galaxies are inhibited by limitations in the available data. The {\\sl ROSAT} \\cite{trump} Position Sensitive Proportional Counter (PSPC) \\cite{pf} has inadequate spatial resolution (PSF $\\sim 30\\arcsec$ FWHM) to map the detailed mass distributions for all but the largest nearby galaxies, and the limited spectral resolution and band width complicates interpretation of the measured temperature profiles (Buote \\& Canizares 1994; Trinchieri et al. 1994; Buote \\& Fabian 1997). Although equipped with superior spatial resolution (PSF $\\sim 4\\arcsec$ FWHM), the {\\sl ROSAT} High Resolution Imager (HRI) \\cite{david} has too small an effective area and too large an internal background to provide images of sufficient quality for many galaxies for radii $r\\ga R_e$. Among the few galaxies with detailed measurements of their radial mass profiles are NGC 507 (Kim \\& Fabbiano 1995), NGC 1399 (Rangarajan et al. 1995; Jones et al. 1997), NGC 4472 (Irwin \\& Sarazin 1996), NGC 4636 (Trinchieri et al. 1994), NGC 4649 \\cite{bm}, and NGC 5044 (David et al. 1994). The shape of the gravitating mass has been measured via X-ray analysis for the E4 galaxy NGC 720 and the E7/S0 galaxy NGC 1332 and found to be at least as elongated as the optical isophotes (Buote \\& Canizares 1994, 1996a, 1997a). For NGC 720, which has more precise constraints, the ellipticity of the gravitating matter is $\\epsilon_{mass}=0.44-0.68$ (90\\% confidence) compared to the intensity weighted ellipticity of the optical light, $\\langle\\epsilon\\rangle=0.31$ (Buote \\& Canizares 1997a). In addition, the X-ray isophotes of NGC 720 twist from being aligned with the optical isophotes within $R_e$ to a position $\\sim 30\\degr$ offset at larger radii. This twist, when combined with the ellipticities of the X-ray isophotes, cannot be explained by the projection of a reasonable triaxial matter distribution and thus may implicate a dark matter halo misaligned from the stars (Buote \\& Canizares 1996b; Romanowsky \\& Kochanek 1997). NGC 720 and NGC 1332 were selected for analysis since they are isolated, significantly elongated in the optical, sufficiently bright, and sufficiently dominated by emission from hot gas in the {\\sl ROSAT} band. In this paper we present X-ray analysis of the classic ``shell'' galaxy, NGC 3923, which is the last galaxy of which we are aware that satisfies these selection criteria and has deep {\\sl ROSAT} observations. This isolated E4 galaxy has both archival {\\sl ROSAT} PSPC and HRI data and its {\\sl ASCA} spectrum has been analyzed previously \\cite{bf}. This will serve as our final case study until the impending launch of {\\sl AXAF} revolutionizes this field. The organization of this paper is as follows. In \\S \\ref{obs} we describe the {\\sl ROSAT} observations and the data reduction. We discuss removal of point sources in \\S \\ref{pt}. Measurements of the ellipticities of the X-ray isophotes and the radial profiles are described in \\S \\ref{e0} and \\S \\ref{radpro} respectively. Analysis of the PSPC spectrum is presented in \\S \\ref{spectra}. We give results for the Geometric Test for dark matter in \\S \\ref{gt} and constraints on the shape and radial mass distribution from detailed hydrostatic models in \\S \\ref{models}. Finally, in \\S \\ref{conc} we give our conclusions. ", "conclusions": "\\label{conc} We have analyzed the gravitating matter distribution of the E4 galaxy NGC 3923 using archival X-ray data from the {\\sl ROSAT} PSPC and HRI. Analysis of the PSPC data, which allows more precise constraints than the HRI data, demonstrates that the X-ray isophotes are significantly elongated with ellipticity $\\epsilon_x=0.15 (0.09-0.21)$ (90\\% confidence) for semi-major axis $a\\sim 10h^{-1}_{70}$ kpc and have position angles aligned with the optical isophotes within the estimated uncertainties. A bright point source located $\\sim 100\\arcsec$ along the major axis inhibits reliable ellipticity constraints for larger radii. By applying a ``Geometric Test'' for dark matter, which essentially compares the shapes of the observed X-ray isophotes to those predicted if mass traces the optical light $L$ (independent of the poorly constrained temperature profile of the gas), we found that the ellipticity of the PSPC X-ray surface brightness exceeds that predicted by the constant $M/L$ hypothesis at the 80\\%-85\\% confidence level. The ``Geometric Test'' result is conservative since it only considers signatures of dark matter that are distributed differently from the optical light. Although the evidence for dark matter from the Geometric Test is marginal, the results from models which employ an explicit solution of the hydrostatic equation assuming an isothermal gas (which is supported by the PSPC spectrum -- \\S \\ref{spectra}) indicate that $M\\propto L$ is highly inconsistent with the radial profiles of the PSPC and HRI data ($\\chi^2_{\\rm red}=3.5$ for 16 dof). This particular discrepancy arises because $L$ is too centrally concentrated: the derived scale length of the gravitating matter is approximately 1.5-2 times that of $L$. The ellipticities predicted by this $M\\propto L$ model fall below the PSPC data at a significance slightly greater than the 90\\% level. Modeling the gravitating mass with a density run $\\rho\\sim r^{-2}$ or with a Hernquist profile we find that the ellipticity of the gravitating matter is, $\\epsilon_{mass}\\cong 0.35 - 0.65$ (90\\% confidence), which is larger than the intensity weighted optical ellipticity $\\langle\\epsilon\\rangle = 0.30$. This evidence for dark matter which is more flattened and more extended than $L$ is similar to our conclusions from previous X-ray studies of two other ellipticals, NGC 720 and NGC 1332, but at somewhat smaller significance level than for NGC 720 (e.g. Buote \\& Canizares 1997b). These results are consistent with analyses of known gravitational lenses (e.g. Keeton, Kochanek, \\& Falco 1997), two polar ring galaxies (Sackett et al. 1994; Sackett \\& Pogge 1995), and flaring disks in spiral galaxies (e.g, Olling 1996). The ellipticities of the gravitating matter derived from our X-ray analyses and these other methods are consistent with those of halos produced by CDM simulations (e.g. Dubinski 1994). If an isothermal gas is assumed then models with matter density $\\rho\\sim r^{-2}$ are favored over Hernquist models (and similar models like the universal CDM profile of Navarro et al. 1997). For $r\\sim 100\\arcsec-300\\arcsec$ the $\\rho\\sim r^{-2}$ model marginally fits the data better than the Hernquist model. However, most of the difference in these models occurs in the central radial bins where the effects of multi-phase cooling flows, magnetic fields, and discrete sources could affect the surface brightness profiles, though we have argued the effects are unlikely to be important (see \\S \\ref{models}). (The derived shape of the gravitating mass is mostly robust to these issues -- Buote \\& Canizares 1997b.) This support for nearly $r^{-2}$ profiles agrees with previous studies of gravitational lenses (e.g. Maoz \\& Rix 1993; Kochanek 1995), although a recent paper finds that density profiles with changing slopes (e.g. Hernquist and NFW) are preferred \\cite{lilya}. An emission component that is proportional to $L$ cannot contribute significantly to the {\\sl ROSAT} X-ray emission of NGC 3923, and thus discrete sources should not affect our constraints on the gravitating matter (Buote \\& Canizares 1997a). However, the {\\sl ASCA} spectral data when fitted with two thermal components yield a cold component, $T_C=0.55$ keV, and a hot component, $T_H\\sim 4$ keV, where the relative flux of cold-to-hot is $\\sim 1.9$ in the {\\sl ROSAT} band \\cite{bf}. The conventional interpretation of the hot component (e.g. Matsumoto et al. 1997; Loewenstein \\& Mushotzky 1997) is that it arises from discrete sources. But our analysis (\\S \\ref{radpro}) shows that $\\sim 35\\%$ of the 0.5-2 keV emission cannot be distributed like the optical light which would be expected of discrete sources. Hence, either the emission from discrete sources is not distributed like $L$, or the hot component obtained from the spectral fits cannot be entirely due to discrete sources as suggested by Buote \\& Fabian \\shortcite{bf}. The constraints we have obtained for NGC 720, NGC 1332, and now NGC 3923 from analyses of their X-ray isophote shapes and radial surface brightness profiles provide an initial demonstration of the power of X-ray analysis for probing the shape and radial distribution of gravitating matter in early-type galaxies. The next generation of X-ray satellites, particularly {\\sl AXAF} and {\\sl XMM}, have the capability to accurately map X-ray isophote shapes and orientations from the cores ($r\\sim 1\\arcsec$) out to 10s of kpc for many galaxies\\footnote{The vastly improved spatial resolution of {\\sl AXAF} over the {\\sl ROSAT} PSPC will allow easy exclusion of the bright point source (1) (see Table \\ref{tab.src}) which hindered the present analysis of NGC 3923.}. The spatially resolved spectra provided by these future missions will allow more precise constraints on temperature gradients and the contribution from discrete sources. Thus, unlike most other methods, obtaining interesting X-ray constraints on the shape and radial density profile of the gravitating matter will be possible for a large sample of early-type galaxies since the X-ray analysis is applicable to any isolated early-type galaxy whose soft X-ray emission ($\\sim 0.5-2$ keV) is dominated by hot gas." }, "9803/astro-ph9803178_arXiv.txt": { "abstract": "A deep, fuly sampled diffraction limited (FWHM $\\sim$ 70 mas) narrow-band image of the central region in M87 was obtained with the Wide Filed and Planetary Camera 2 of the {\\it Hubble Space Telescope} using the dithering technique. The \\HaNii\\ continuum subtracted image reveals a wealth of details in the gaseous disk structure described earlier by Ford et al.\\ (1994). The disk morphology is dominated by a well defined three-arm spiral pattern. In addition, the major spiral arms contain a large number of small ``arclets'' covering a range of sizes (0\\as1--0\\as3 = 10--30 pc). The overall surface brightness profile inside a radius $\\sim$ 1\\farcs5 (100 pc) is well represented by a power-law $I(\\mu) \\sim \\mu^{-1.75}$, but when the central $\\sim$ 40 pc are excluded it can be equally well fit by an exponential disk. The major axis position angle remains constant at about PA$_{\\rm disk} \\sim 6^{\\circ}$ for the innermost $\\sim 1''$, implying the disk is oriented nearly perpendicular to the synchrotron jet (PA$_{\\rm jet} \\sim 291^{\\circ}$). At larger radial distances the isophotes twist, reflecting the gas distribution in the filaments connecting to the disk outskirts. The ellipticity within the same radial range is $e = 0.2-0.4$, which implies an inclination angle of $i \\sim 35^{\\circ}$. The sense of rotation combined with the dust obscuration pattern indicate that the spiral arms are trailing. ", "introduction": "The disk of ionized gas in the nucleus of M87 is currently the best example of a family of similar small ($r \\sim 100$ pc) gaseous disks found to be common in the centers of elliptical galaxies with active nuclei (for a review see Ford et al.\\ 1998). Several \\HST\\ kinematical studies have shown that in M87 the gas is in Keplerian rotation, orbiting a massive black hole with a mass $M_{\\rm BH} \\sim 2 - 3 \\times 10^9 M_{\\odot}$ (Harms et al.\\ 1994; hereafter H94, Ford et al.\\ 1996a,b; and Macchetto et al.\\ 1997, hereafter M97). The few other galaxies studied kinematically so far (NGC 4261 -- Ferrarese et al.\\ 1996, NGC 6521 -- Ferrrarese et al.\\ 1998, NGC 4374 -- Bower et al.\\ 1998) have further shown that nuclear gaseous disks offer an excellent tool for measuring the central black hole mass. Recent studies have revealed other important characteristics of the nuclear disk in M87. Its aparent minor axis (F96, M97) is closely aligned with the synchrotron jet ($\\Delta\\theta \\sim 10^{\\circ} - 15^{\\circ}$) suggesting a causal relationship between the disk and the jet. The system of filaments in the center of M87 (Sparks, Ford \\& Kinney 1993; SFK) may also be causally connected to the disk. For example, the filaments extending $\\sim$17$''$ (1200 pc) to the NW at PA $\\sim 315^{\\circ}$ are blue shifted with respect to systemic velocity and show dust absorption implying they are on the near side of M87 as is the jet. These two findings led SFK to conclude that these filaments are streamers of gas flowing away from the center of M87 rather then falling into it. The images in F94 (see also Ford \\& Tsvetanov, this volume, FT98) show an apparent connection between at least some of the larger scale fillaments and the ionized nuclear disk. Direct spectroscopic evidence for an outflow was found recently. Several UV and optical absorption lines from neutral and very mildly ionized gas were measured in the FOS spectrum of the nucleus (Tsvetanov et al.\\ 1998; T98). These lines are broad (FWHM $\\sim 400$ \\kms) and blue shifted by $\\sim 150$ \\kms\\ with respect to M87's systemic velocity implying both an outflow and turbulence. In addition, non-circular velocity components -- both blue and red shifted -- were found at several locations in the disk (F96, FT98), and observed emission lines are much broader than the expected broadening due to the Keplerian motion accross the FOS aperture. All these properties are best understood if a bi-directional wind from the disk were present. This wind may be an important mechanism for removing angular momentum from the disk to allow accretion through the disk onto the central black hole. Whatever the physical conditions in the disk it is important to map its morphology in detail. The first \\HST\\ images (F94) have hinted that a spiral pattern could be present, but the signal-to-noise was too low for a definitive conclusion. In this paper we present deep, fully sampled diffraction limited narrow band images of the nuclear region in M87. We use these images to characterize the ellipticity, brightness distribution, and morphology of the disk. In this paper we adopt a distance to M87 of 15 Mpc, corresponding to a scale of 1$''$ = 73 pc. \\vspace{-2mm} ", "conclusions": "" }, "9803/astro-ph9803099_arXiv.txt": { "abstract": "We present a catalog of 200 clusters of galaxies serendipitously detected in 647 \\ROSAT\\/ PSPC high Galactic latitude pointings covering 158 square degrees. This is one of the largest X-ray selected cluster samples, comparable in size only to the \\ROSAT\\/ All-Sky Survey sample of nearby clusters (Ebeling et al.\\ 1997). We detect clusters in the inner 17.5\\arcmin\\ of the \\ROSAT\\/ PSPC field of view using the spatial extent of their X-ray emission. Fluxes of detected clusters range from $1.6\\times10^{-14}$ to $8\\times10^{-12}\\,$\\ergs\\ in the 0.5--2~keV energy band. X-ray luminosities range from $10^{42}~$erg~s$^{-1}$, corresponding to very poor groups, to $\\sim5\\times10^{44}\\,$erg~s$^{-1}$, corresponding to rich clusters. The cluster redshifts range from $z=0.015$ to $z>0.5$. The catalog lists X-ray fluxes, core-radii, spectroscopic redshifts for 73 clusters and photometric redshifts for the remainder. Our detection method, optimized for finding extended sources in the presence of source confusion, is described in detail. Selection effects necessary for a statistical analysis of the cluster sample are comprehensively studied by Monte-Carlo simulations. We have optically confirmed 200 of 223 X-ray sources as clusters of galaxies. Of the remaining 23 sources, 18 are likely false detections arising from blends of unresolved point X-ray sources, and for 5 we have not obtained deep CCD images. Above a flux of $2\\times10^{-13}\\,$\\ergs, 98\\% of extended X-ray sources are optically confirmed clusters. The $\\log N - \\log S$ relation for clusters derived from our catalog shows excellent agreement with counts of bright clusters derived from the \\emph{Einstein}\\/ Extended Medium Sensitivity Survey (Henry et al.\\ 1992) and \\ROSAT\\/ All-Sky Survey (Ebeling et al.\\ 1997). At fainter fluxes, our $\\log N - \\log S$ relation agrees with the smaller-area WARPS survey (Jones et al.\\ 1998). Our cluster counts appear to be systematically higher than those from a 50~deg$^2$ survey of Rosati et al.\\ (1998). In particular, at a flux of $2\\times10^{-13}\\,$\\ergs, we find a surface density of clusters of $0.57\\pm0.07$ per square degree, which is a factor of 1.3 more than found by Rosati et al. This difference is marginally significant at the $\\sim 2$ sigma level. The large area of our survey makes it possible to study the evolution of the X-ray luminosity function in the high luminosity range inaccessible with other, smaller area \\ROSAT\\/ surveys. ", "introduction": "Clusters of galaxies are among the most important objects for cosmological studies. Models of large scale structure formation such as CDM, predict that the abundance of clusters is determined by the spectrum of primordial perturbations and cosmological parameters $\\Omega$ and $\\Lambda$. Observations of clusters at different redshifts can be used to constrain these parameters (e.g., White \\& Rees 1978, Kaiser 1986, White, Efstathiou, \\& Frenk 1993, Henry \\& Arnaud 1991, Viana \\& Liddle 1996, Henry 1997). Following a different approach, observations of the Sunyaev-Zel'dovich effect (Sunyaev \\& Zel'dovich 1972) in a large sample of distant clusters can be used for a direct measurement of the distance to these clusters, and thus provide the values of $H_0$ (e.g., Birkinshaw, Hughes, \\& Arnaud 1991) and~$q_0$. Up until the present, the largest samples of distant clusters resulted from optical surveys that searched for enhancements in the surface density of galaxies (e.g., Postman et al.\\ 1996). This method suffers seriously from projection effects (e.g., van Haarlem et al.\\ 1997). Distant clusters found by such techniques as galaxy concentrations around distant radio sources (Dickinson 1996) or ``dark'' lenses (Hattori et al.\\ 1997) cannot be considered as statistical samples. Of all methods for detecting distant clusters, X-ray surveys are the least sensitive to projection, because the X-ray emission is proportional to the square of the density of the hot gas, which must be compressed in a deep potential well for us to detect it. It is noteworthy that unlike optical, X-ray surveys have the possibility of finding interesting objects such as ``fossil'' clusters in which almost all galaxies have merged to form a cD galaxy (Ponman et al.\\ 1994), and hypothetical ``failed'' clusters in which galaxy formation was suppressed (Tucker et al.\\ 1995). To date, the largest published sample of distant X-ray selected clusters is that from the \\emph{Einstein}\\/ Extended Medium Sensitivity Survey (EMSS; Goia et al.\\ 1990, Stocke et al.\\ 1991). However, because of the relatively high flux limit, the EMSS sample contains only 6 clusters at $z>0.5$. Finding clusters in X-rays is complicated by their rarity among other types of sources. A comparison of the $\\log N - \\log S$ relations for all sources (Hasinger et al.\\ 1993a) and clusters (this work) shows that at a flux of $10^{-14}\\,$\\ergs\\ in the 0.5--2~keV band, clusters comprise not more than 10--20\\% of the total source population. The large amount of optical identification work needed for cluster selection can be greatly reduced if they are searched for among spatially extended X-ray sources. Even at $z=1$, a rich cluster with a core-radius of 250~kpc has an angular radius of $>20\\arcsec$, which still can be resolved with the \\ROSAT\\/ PSPC on-axis. Detection of extended sources requires new analysis techniques. Even if the spatial extent is not used for cluster selection, special detection techniques are needed because clusters at $z\\approx 0.2-0.3$ are 3--4 times broader than the \\ROSAT\\/ PSPC point spread function. The idea of selecting distant cluster samples from various \\ROSAT\\/ surveys was pursued by different groups in the past few years. Rosati et al.\\ (1995, 1998) searched for clusters in long exposure ($>15$~ksec) \\ROSAT\\/ PSPC pointed observations with a total area of 50~deg$^{2}$, using optical identifications of all extended X-ray sources found by wavelet transform analysis. Their sample consists at present of 70 clusters. The Wide Angle \\ROSAT\\/ Pointed Survey (WARPS, Scharf et al.\\ 1997, Jones et al.\\ 1998) uses the Voronoi Tessellation and Percolation technique to detect both point-like and extended sources, followed by optical identifications of all sources. The WARPS cluster sample consists at present of 46 clusters found in \\ROSAT\\/ pointings with exposures $>8$~ksec, covering 16.2~deg$^{2}$. A small sample of 15 clusters at $0.3 5000$), homogenous set of spectra. We have shown that we can achieve classification errors of \\sig68\\ = 0.82 subtypes (\\sigrms\\ = 1.09 subtypes) over this complete range of spectral subtypes. This result compares favourably with the intrinsic errors of \\sig68 = 0.63 subtypes in our training data. Once a neural network has been trained, its classification results are completely reproducible. Moreover, the low values of their internal errors ($<0.4$ spectral subtypes) demonstrate that networks can be re-trained to give sufficiently consistent classifications. We have achieved correct luminosity class classification for over 95\\% of dwarfs (class V) and giants (class III). Results for luminosity class IV spectra were considerably worse. It is believed that the data themselves could be a limiting factor and methods for improving these results were discussed. Despite the correlation in the data set between spectral type and luminosity class, it was demonstrated that the neural networks were using luminosity features to do dwarf-giant discrimination. Network with two hidden layers performed considerably better ($\\approx 0.2$ subtypes) than ones with only one hidden layer. The best classification results were achieved by tackling the spectral type and luminosity class problems separately, using continuous and probabilistic networks respectively. We used Principal Components Analysis to compress the spectra by a factor of over 30 while retaining 96\\% of the variance in the data. It was shown that this compression predominantly removes noise. In addition the PCA preprocessing reduces the dimensionality of the data and can be used to filter out bogus spectral features or identify unusual spectra. However, PCA has the drawback that very weak or rare features will not be well-reconstructed. More complex non-linear preprocessing schemes could no doubt be devised, but the strength of PCA is its analytic simplicity and its robustness. The automated classifiers presented in this paper have been used to produce classifications for several thousand stars which do not have classifications listed in the MHD catallogue. These will be presented in a future paper (Bailer-Jones 1998)." }, "9803/astro-ph9803320_arXiv.txt": { "abstract": "We have made a series of joint spectral fits for two blank fields, the Lockman Hole and the Lynx-3A field, where a significant amount of both {\\it ASCA} and {\\it ROSAT} PSPC data exist after thorough screenings. The {\\it ASCA} SIS, GIS and {\\it ROSAT} PSPC spectra from these fields have been fitted simultaneously. Comparison at $E>1$ keV shows general agreement within 10\\% in the Lockman Hole data and a $20-30\\%$ disagreement in the Lynx-3A data, indicating remaining observation-dependent systematic problems. In both cases, satisfactory fits have been found for the overall 0.1-10 keV spectrum with an extragalactic power-law component (or a broken power-law component with steepening at $E<1$ keV), a hard thermal component with plasma temperature of $kT^{\\rm h}\\approx0.14$ keV and a soft thermal component $kT^{\\rm s}\\approx0.07$ keV. ", "introduction": "\\label{sec:intr} The global spectrum of the cosmic X-ray background is a primary piece of information for understanding its origin. The 3-50 keV CXRB spectrum observed with HEAO-1 A2 can be well described by a $kT=40$ keV thin thermal plasma-like spectral shape (Marshall et al. \\cite{marshall80}; Boldt \\cite{boldt87}), which can be approximated by a power-law with a photon index of $\\Gamma = 1.4$ in $E\\la 10 keV$. {\\it BBXRT} (Jahoda et al. \\cite{jahoda92}) and {\\it ASCA} (Gendreau et al. \\cite{gend_spec}; Ishisaki et al. \\cite{ishi98}) measurements show that this power-law component extends down to 1 keV, below which an excess is observed. A number of authors report {\\it ROSAT} measurements in the 0.5-2 keV band (e.g. Hasinger \\cite{has92}; Georgantopoulos et al. \\cite{georg96}) and show about 30\\% larger flux than the Gendreau et al.'s (\\cite{gend_spec}) {\\it ASCA} SIS result at 1 keV, with different slopes (see Hasinger \\cite{has96} for review). The disagreement may be contributed by the differences in the position/solid angle of the measured sky, problems arising from incomplete modelings, and/or calibration problems. In order to separate these effects, we have made a series of joint spectral fits of {\\it ROSAT} PSPC, {\\it ASCA} GIS, and {\\it ASCA} SIS spectra from two fields of the sky, where sufficient amount of blank-sky data exist after thorough screening. Because of the limited data meeting the criteria, the work presented in this paper is not intended to determine the current best estimate of the global CXRB spectrum, but rather a comparison of measurements among {\\it ASCA} and {\\it ROSAT} instruments in the same parts of the sky with consistent modelings. In Sect.\\ref{sec:data}, we describe the {\\it ASCA} and {\\it ROSAT} data used in the analysis. Joint spectral fits are described in Sect.\\ref{sec:fit}. The results are discussed in Sect.\\ref{sec:disc}. ", "conclusions": "\\label{sec:disc} There is a bright variable source in LH with [1.2$\\pm$.2] and [2.5$\\pm$.7] $\\times 10^{-13} [{\\rm erg\\,cm^{-2}\\,s^{-1}}]$ for the 0.7-2 keV and 2-7 keV bands respectively (Ogasaka \\cite{oga_t}), consisting about 10\\% of the total ASCA fluxes in both bands. This source was much fainter during the PSPC observation. Thus one should decrease the GIS and SIS normalizations about $10\\%$ lower when comparing with the PSPC data. In this case, the agreement between PSPC and GIS is excellent and falls well within statistical errors of each other. For both LH and LX, the SIS data consistently show $\\approx 10\\%$ lower normalizations compared to GIS. This might be caused by incomplete calibration for the radiation damage of the SIS with the 4CCD mode, which can even exist at this level after a few months after the launch, when LH and LX were observed (Dotani et al. \\cite{dotani95}). In the LX observation, a larger discrepancy exists. The GIS and SIS normalizations are lower than the PSPC value by $\\approx20\\%$ and $\\approx30\\%$ respectively and slopes are shallower. The ASCA LX normalizations are also significantly lower than those of LH. There is no variable source which can cause this amount of discrepancy in LX. The fact that the 0.1-10 keV fit still show the disagreement of the normalization (see $N_{\\rm G}$ in B2) shows that this is not a modeling problem (e.g. leak of the $E>1$ keV excess with the PSPC energy resolution). One possible explanation is the LTE (e.g Snowden et al. \\cite{snow94}), which is usually apparent in the $E<0.5$ keV channels of the PSPC data, but sometimes extends above 1 keV when the activity is high. There may also be an over-subtraction of the NXB background from the {\\it ASCA} data. Furthermore, the instruments are not looking at exactly the same part of the sky. Due to the stray-light and PSF of the ASCA instruments, $\\approx 40\\%$ of the GIS/SIS flux comes from outside of the designated FOV (estimated using our ray-tracing program). These effect may also contribute to this discrepancy. \\begin{figure}[t] \\psfig{file=lo_comp2.ps,width=\\hsize,angle=270} \\caption[]{The PSPC and GIS $E\\,I(E)$ spectra (using a two power-law model as an appropriate smooth function for unfolding purpose) of LH are shown and compared with previous measurements: the thick solid bowtie is from Hasinger (\\cite{has92}); the dot-dashed bowtie from Georgantopoulos et al. (\\cite{georg96}), both used {\\it ROSAT} PSPC. The dotted line is from rocket measurements (McCammon \\& Sanders \\cite{maccamon90}). The long-dashed horn is from an {\\it ASCA} SIS measurement by Gendreau et al. \\cite{gend_spec} and the thin solid bowtie is a joint {\\it ROSAT} PSPC/{\\it ASCA} SIS analysis of QSF3 by Chen et al. (\\cite{chen97}) for $E>1$ keV. The thick solid line represents the HEAO-1 A2 measurement by Marshall et al. (\\cite{marshall80}).} \\label{fig:lo_comp} \\end{figure} The best consistency for a certain region of the sky from this work is seen for the PSPC and GIS data on LH. Thus it is instructive to compare the observed spectra of these with previous CXRB measurements. The comparison is shown in Fig. \\ref{fig:lo_comp}. Fig. \\ref{fig:lo_comp} shows a large excess at $E\\sim 0.6$ keV on the PSPC data, inconsistent with Gendreau et al.'s {\\it ASCA} SIS data. This may be partially due to the low Galactic column density of LH. The LH GIS data for $E\\ga 2$ keV are above the HEAO-1 A2 and Gendreau et al. (\\cite{gend_spec}) SIS values. We note, however, that about 10 \\% of source fluctuation is expected over this small area. Since the ASCA LH data contains a bright source ($\\sim 5\\times 10^{-13} {\\rm erg\\,s^{-1}\\,cm^{-2}}$ in 2-10 keV), this field should be one of the brighter ones. We also note, however, that an integration from the brightest source in the field to the faintest source excluded in the collimator experiments (e.g. $\\approx2\\times 10^{-11} {\\rm erg\\,s^{-1}\\,cm^{-2}}$ in 2-10 keV for HEAO-1 A2 measurement by Marshall et al.) would add $\\approx10\\%$ of intensity. A thorough treatment of source fluctuations using a larger area and comparing spectra with appropriate source removal will be presented in a future paper. In summary, a close look at {\\it ROSAT} and {\\it ASCA} spectra for the same regions of the sky have revealed systematic errors caused by response calibration problems and non cosmic background subtraction of up to $\\approx 20-30\\%$ for one set of observations. These probably caused the reported disagreements between {\\it ASCA} and {\\it ROSAT} measurements (Hasinger 1996), while modelings and sky selection can also contribute. We have obtained a fair description of the CXRB spectrum over 0.1-10 keV range cosisting of a extragalactic power-law component (either single or broken below 1 keV), hard and soft thermal components with a satisfactory fit to all instruments." }, "9803/astro-ph9803116_arXiv.txt": { "abstract": " ", "introduction": "Session B.3 received a partisan organisation, and was divided into sections corresponding to the main paradigms pervading modern cosmology. Three sub-sessions were allocated to cover inflationary cosmology, pre-big-bang scenarios, and topological defects in cosmology. Anything not fitting into these topics makes up the last section, covering miscellaneous topics. Below I briefly review the current status in each subject covered in a sub-session after which I summarise the talks presented. These summaries reflect my personal understanding of the talks, and I apologise to the speakers if I accidentally missed the entire point. ", "conclusions": "" }, "9803/astro-ph9803184_arXiv.txt": { "abstract": "Blue- and red-shifted Hydrogen and Helium satellite recombination lines have recently been discovered in the optical spectra of at least two supersoft X-ray sources (SSSs), RX~J0513-069 and RX~J0019.8+2156, and tentatively also in one short-period cataclysmic variable star (CV), the recurrent nova T~Pyx. These features are thought to provide evidence for the presence of highly collimated jets in these systems. No similar spectral signatures have been detected in the spectra of other short-period cataclysmic variables, despite a wealth of existing optical data on these systems. Here, we ask if this apparent absence of ``jet lines'' in the spectra of most CVs already implies the absence of jets of the kind that appear to be present in the SSSs and perhaps T~Pyx, or whether the current lack of jet detections in CVs can still be ascribed to observational difficulties. To answer this question, we derive a simple, approximate scaling relation between the expected equivalent widths of the observed jet lines in both types of systems and the accretion rate through the disk, $EW(line) \\propto \\dot{M}_{acc}^{\\frac{4}{3}}$. We use this relation to predict the strength of jet lines in the spectra of ``ordinary'' CVs, i.e. systems characterized by somewhat lower accretion rates than T~Pyx. Making the assumption that the features seen in T~Pyx are indeed jet lines and using this system as a reference point, we find that if jets are present in many CVs, they may be expected to produce optical satellite recombination lines with EWs of a few hundredths to a few tenths of Angstroms in suitably selected systems. A similar prediction is obtained if the SSS RX~J0513-069 is used as a reference point. Such equivalent widths are small enough to account for the non-detection of jet features in CVs to date, but large enough to allow them to be detected in data of sufficiently high quality, if they exist. ", "introduction": "\\label{introduction} One of the most intruiging empirical connections that has emerged from observations of accretion-powered objects on all astrophysical scales is that between accretion disks and powerful bipolar outflows and jets. Indeed, the occurrence of mass loss in this form has been established in members of essentially all classes of (presumed) disk-accretors, including close binary systems such as X-ray binaries, supersoft X-ray sources (SSSs) and non-magnetic cataclysmic variable stars (CVs), young stellar objects such as T-Tauri and FU~Orionis stars, and active galactic nuclei and quasars. However, a complete theoretical understanding of this empirical disk-wind (or disk-jet) connection has not yet been achieved. Recent reviews of the subject with references to the relevant observational literature may be found in Livio \\shortcite{livio4,livio5}. A potentially vital clue to the origin of mass loss from accretion disk systems is provided by the fact that in at least some members of all but one class of objects the outflow collimation appears to be very tight, with half-opening angles, $\\theta_{max}$, of no more than a few degrees (we will refer to such flows as {\\em jets} hereafter). The sole exception to this rule are CVs (see however below). In these systems, the presence of mass loss has nevertheless been clearly established, based on the shapes and eclipse behavior of the ultraviolet (UV) resonance lines (e.g. Drew 1991), but the inferred collimation of the corresponding outflows is weak ($\\theta_{max} \\gtappeq 45^o$; Shlosman, Vitello \\& Mauche 1997; Knigge \\& Drew 1997). The apparent absence of jets in CVs may hold the key to an improved theoretical understanding of the disk-wind connection. If confirmed, any ``generic'' model (in the sense of being applicable to more than a particular class of objects) for the origin of mass loss from accretion disks must be able to explain why this mass loss takes the form of jets in all systems but CVs. This would be quite a restrictive constraint, particularly in the light of two recent observational developments. The first of these is the detection of blue- and red-shifted satellite emission features to the optical Hydrogen and Helium recombination lines in the SSSs RX~J0513-069, RX~J0019.8+2156 and (possibly) CAL~83 \\cite{crampton1,crampton2,southwell1,southwell2,becker1}. These features cannot be attributed to any other ionic species and are therefore thought to arise in some kind of bipolar outflow. The combination of small width and large displacement from line center exhibited by these satellite lines further suggests that the corresponding outflows are very highly collimated, i.e. that they are jets (c.f. the prototypical jet system SS~433; Vermeulen~1993). If this identification is correct (which we will assume throughout this paper) it is significant, because SSSs and CVs are extremely similar types of objects: both are semi-detached binary systems in which a Roche-lobe filling secondary star transfers material via an accretion disk onto a white dwarf (WD) primary. The main difference between SSSs and CVs is thought to be the rate at which this mass transfer proceeds. In SSSs, the accretion rate is believed to be high enough ($\\dot{M}_{acc} \\sim 10^{-7} - 10^{-6}$~M$_{\\sun}$~yr$^{-1}$) to fuel steady nuclear hydrogen burning on the surface of the WD (e.g. van den Heuvel et a. 1992). By contrast, the highest accretion rates encountered in CVs are about one order of magnitude lower than this ($\\dot{M}_{acc} \\sim 10^{-8}$~M$_{\\sun}$~yr$^{-1}$) and thus insufficient to initiate steady nuclear burning on the WD. Instead, the accretion can result in shell flashes which are responsible for nova outbursts. This difference is one of the factors that led Livio (1997a) to propose that the formation of powerful jets (as opposed to more weakly collimated bipolar outflows) may {\\em require} the presence of an additional wind/energy source at the center of the accretion disk. The second recent observational finding of significance in the present context is the tentative detection of similar H$\\alpha$~satellite lines in the optical spectrum of {\\em one} CV, the recurrent nova T~{Pyx} \\cite{shahbaz1}. It is important to stress that no similar features have ever been detected in any other short-period CV, despite the fact that optical spectra of high enough quality to detect satellite lines of similar strength as in T~Pyx (which, on average, have an equivalent width of about 1~\\AA) should be available for many of them. While it is possible that in some cases they might have been overlooked (previous studies would not have expected to see such features), it would nevertheless appear that satellite lines of the strength seen in T~Pyx are not a common feature among CVs. For example, Shahbaz et al. (1997) did not detect similar satellite lines in the spectrum of another recurrent nova, U~Sco. (It should be noted that one-sided satellite lines have been seen in the spectra of a few CVs, e.g. S193, V795~Her, BT~Mon [Szkody 1995; Haswell et al. 1994; Seitter 1984]; however, these probably arise in other types of high velocity flows in these systems.) The physical similarities between CVs and SSSs, on the one hand, and the observational similarity between the satellite features in T~Pyx and those in SSSs such as RX~J0513-069, on the other, suggest that T~Pyx may also harbor a well collimated jet. A fundamental assumption we adopt in the present paper is that this is indeed the case. Note that, under this assumption, we would expect T~Pyx's binary inclination to be higher than that of the SSSs listed above, since in T~Pyx the ratio of satellite line widths to their displacements from line center is somewhat smaller. A high inclination for T~Pyx ($i \\sim 70^o$) is also indicated if one demands that the displacement of the satellite features from line center should roughly correspond to the escape velocity from the WD, as expected if they are formed in a jet (c.f. Shahbaz et al. 1997; Livio 1997a). It is acknowledged, however, that Shahbaz et al. (1997) also derived an inclination estimate from the peak-to-peak separation of the double-peaked H$\\alpha$ line core, which, by contrast, turns out to be very low ($i \\sim 10^o$). While this estimate should be regarded as a lower limit \\cite{shahbaz1}, the case for a well collimated jet in T~Pyx would have to be critically reexamined if the inclination of this system were shown to be $\\ltappeq 50^o$ in the future. The analytic scaling relation derived in Section~2 would of course nevertheless be valid and, we believe, useful, even if T~Pyx should eventually turn out not to contain a collimated jet. Actually, the existence of a jet in T~Pyx is not entirely unexpected, since it is thought that recurrent novae in general, and T~Pyx in particular, are characterized by accretion rates that are higher (i.e. $\\dot{M}_{acc} \\gtappeq 10^{-8}$~M$_{\\sun}$~yr$^{-1}$) than those in other types of CVs, such as dwarf-novae (DNe) and nova-like variables (NLs). In fact, Webbink et al. (1987) have suggested that intermittent nuclear burning might take place on the surface of the WD in T~Pyx even during quiescence, i.e. between nova outbursts. If so, Livio's (1997a) hypothesis could be used to reconcile the presence of jet lines in SSSs and T~Pyx with the absence of similar features in the spectra of other CVs. The main goal of the present paper is to determine whether such reconciliation is actually required at present. Thus, we will ask whether the apparent absence of jet lines in the existing optical spectra of most CVs, particularly NLs and DNe in outburst (i.e. systems in which an optically thick accretion disk is present but no nuclear burning occurs) already implies the {\\em absence} of jets of the kind seen in the SSSs and (possibly) T~Pyx, or whether the current lack of jet detections in CVs can still be ascribed to observational difficulties. In our attempt to answer this question, we will take as our fundamental working hypothesis that the main difference between CVs and the SSSs (as well as perhaps T~Pyx) -- the presence of an extremely hot WD at disk center -- is irrelevant to the formation of the observed jets. The corollary of this hypothesis is that CVs must actually drive jets of the same kind that appear to be present in the SSS and T~Pyx. Our goal is to estimate the expected equivalent widths of CV jet lines within the framework of this hypothesis. In so doing, we will try to make as few references as possible to specific models for the jet formation and line emission mechanisms, and instead use only simple and general physical arguments. As already noted above, under our working hypothesis, jets akin to those in the SSSs must actually be present in CVs. This is not in direct conflict with the relatively wide outflow opening angles that have been inferred for CV winds from modeling the UV resonance lines: these spectral features probe only the near-disk regime of the outflow -- out to at most a few hundred $R_{WD}$ -- leaving collimation at larger distances as a distinct possibility. ", "conclusions": "\\label{BLA} \\subsection{A prediction for the strength of jet lines in CVs} \\label{BLA2} We are now in a position to use Relation~(\\ref{final2}) to predict the jet line EWs we would expect to see in ``ordinary'' CVs, according to our working hypothesis. Since T~Pyx {\\em is} in fact a CV, its system parameters are more typical of ``normal'' CVs than are those of the SSSs exhibiting jet lines. It is therefore preferable to use T~Pyx as a reference point in making predictions for other CVs, because the ratios of the relevant factors in Relation~(\\ref{final2}) will be closer to unity and the associated uncertainties will be smaller. However, in Section~\\ref{BLA3} below we will check whether Relation~(\\ref{final2}) is at least consistent with the observed accretion rate and EW ratios of T~Pyx and the SSS RX~J0513-069. (See also the note at the end of the manuscript, in which we show that the prediction derived in this section by using T~Pyx as a reference datum is consistent with what is be obtained if RX~J0513-069 is used instead.) Concerning T~Pyx, Webbink et al. (1987) give $\\dot{M}_{acc}(T~Pyx) \\gtappeq 10^{-8}$~M$_{\\sun}$~yr$^{-1}$ based on the short recurrence time scale of its eruptions and $\\dot{M}_{acc}(T~Pyx) \\sim 5 \\times 10^{-8}$~M$_{\\sun}$~yr$^{-1}$ based on its optical colors. Here, we adopt the latter estimate, which assumes that the optical light is due to the accretion disk, rather than to direct and/or reprocessed light from a hot WD. This is in line with our working hypothesis that a wind/energy source at disk center is not required to drive jets from accretion disks. T~Pyx's orbital period is thought to be $P_{orb}(T~Pyx) \\simeq 1.8$~hrs \\cite{schaefer1}, placing the system below the period gap. As noted in Section~\\ref{introduction}, the inclination angle of T~Pyx is not well constrained observationally, although the appearance of the satellite recombination lines themselves suggests a high value $i(T~Pyx) \\simeq 70^o$ if these features are formed in a well collimated jet. Finally, the equivalent widths of the H$\\alpha$~jet lines in T~Pyx can be measured from the data of Shahbaz et al. (1997) and turn out to be about EW(T~Pyx) $\\simeq 1$~\\AA~on average, with the strongest feature in any one of the observing epochs reaching about twice this value (Shahbaz, private communication). We now need to make some assumptions about the typical properties of ``ordinary'' CVs. The form of Relation~(\\ref{final2}) shows that if jets are present in these objects, the associated satellite recombination lines are likely to be strongest in systems with high mass accretion rates, short orbital periods and high inclinations (though not so high as to shift the jet lines into the line core). Since it would be sufficient to falsify our working hypothesis for CVs with these properties, we adopt $\\dot{M}_{acc}(CV) \\simeq 1 \\times 10^{-8}$~M$_{\\sun}$~yr$^{-1}$ (appropriate to NLs and DNe in outburst), $P_{orb}(CV) \\simeq P_{orb}(T~Pyx)$, and $i(CV) \\simeq i(T~Pyx)$. (We note in passing that there are actually no well-established non-magnetic NL variables with periods shorter than 3.2~hrs, although there is a fair number of DNe with $P_{orb} \\leq 2$~hrs.) We can now use Relation~(\\ref{final2}) to predict the jet line EWs we expect to see in this most favorable sub-group of ``ordinary'' CVs, according to our working hypothesis. To this end, we take the ratio of the two separate relations (one for the normal CVs, one for T~Pyx), solve for $EW(CV)$ and substitute our adopted parameters. This yields $EW(CV) \\simeq 0.1^{+0.1}_{-0.04}$~\\AA. While it may be possible to increase the upper limit implied by this result somewhat -- by taking $P_{orb}(CV)$ to be shorter or $i(CV)$ to be higher, for example -- it is clear that CV jet lines, if they exist, would be at best marginally detectable in typical optical spectra. As a result, we are forced to conclude that our working hypothesis and its corollary -- that a hot central object is inessential to the formation of jets and that CVs do in fact drive jets -- {\\em cannot} yet be ruled out. Let us take a step back at this point to make it clear what we are -- and are not -- claiming. We started by adopting the working hypothesis that the formation of powerful jets does {\\em not} require the presence of an additional energy source at disk center. As a corollary, we assumed that ``ordinary'' CVs harbor the same kind of jets that may be present in T~Pyx (as indicated by the satellite recombination lines that are observed in that object). We then showed that based on these assumptions one can derive a simple scaling law which can be used to predict the expected strength of these jet lines in the optical spectra of ordinary CVs. The predicted jet line EWs for these systems turned out be very small, even for objects with nearly optimal system parameters. We therefore concluded that the lack of jet line detections in the optical spectra of ordinary CVs is not yet in conflict with our working hypothesis, i.e. that jets {\\em may} be present in ordinary CVs. Note that we do {\\em not} claim to have shown that ordinary CVs actually {\\em do} contain jets. After all, an inability to falsify a hypothesis does not prove it. Summarized succinctly, our conclusion is that {\\em the non-detection of jet lines in existing optical spectra of ``ordinary'' CVs should not yet be taken to imply that these systems cannot harbor collimated jets}. Two further points need to be made regarding this statement. First, even though we have been unable to rule out the presence of jets in ``ordinary'' CVs on the basis of existing data, the predicted EWs of a few hundredths up to a few tenths of Angstroms may not be beyond the reach of high resolution, high signal-to-noise optical spectra. Thus we strongly encourage observers to search for the signatures of jets in the spectra of appropriately selected CVs. Second, it was assumed above that T~Pyx's optical continuum is dominated by the radiation field emitted by a standard accretion disk. However, if (intermittent) nuclear burning really does take place in T~Pyx, the surface of the WD at the center of the disk will be extremely hot. It is therefore worth asking whether (some of) the optical continuum could actually be direct or reprocessed radiation emitted by the WD, and what effect this may have on our conclusions. A numerical calculation similar to that described following Relation~(\\ref{continuum1}) in Section~\\ref{scale} shows that direct light from the WD is unlikely to be of any importance, even if the temperature of the WD is as high as a few times $10^5$~K, and the accretion rates as low as $10^{-8}$~M$_{\\sun}$~yr$^{-1}$. To judge the potential significance of reprocessed WD radiation, we rely on the recent work of King (1997), who derived a simple condition that can be used to estimate the relative importance of dissipation and reprocessing in a CV accretion disk. More specifically, King (1997) showed that reprocessing of WD radiation will begin to have a dominant effect on the local disk temperature if $L_{WD} \\gtappeq 2.5 L_{acc} (1-\\beta)^{-1}$, where $L_{WD}=4\\pi R_{WD}^2 \\sigma T_{WD}^4$ and $L_{acc} = GM_{WD} \\dot{M}_{acc}/R_{WD}$ are the WD and total accretion luminosities, respectively, and $\\beta$ is the albedo of the disk surface. To give a numerical example, we note that if reprocessing is assumed to be efficient ($\\beta \\simeq 0$), the temperature distribution in a disk around a $1M_{\\sun}$~WD accreting at a rate of $\\dot{M}_{acc} = 10^{-8} M_{\\sun}$~yr$^{-1}$ will be dominated by reprocessing if $T_{WD} \\gtappeq 2 \\times 10^5$~K. If reprocessed WD radiation is in fact contributing significantly to T~Pyx's optical continuum, then our previous prediction for the strength of jet lines in other CVs no longer applies, since our continuum scaling law, Relation~(\\ref{continuum}), ceases to be valid. Qualitatively, the effect of this will be to increase the predicted EWs significantly, since (a) the adopted accretion rate for T~Pyx is almost certainly an overestimate in this case, and (b) the extra contribution to the continuum that is ultimately due to nuclear burning on the WD (and not to accretion) is making the jet lines appear weaker than if only the disk were producing the continuum. Quantitatively, these effects can be corrected for by multiplying the predicted EWs by a factor of $f_{\\dot{M}}^{2}$, where $f_{\\dot{M}}>1$ is the factor by which T~Pyx's accretion rate has been overestimated. The dependence on $f_{\\dot{M}}$ {\\em squared} arises because the part of correction (a) that is related to the scaling of the continuum flux with accretion rate exactly cancels correction (b). This leaves the scaling of the line luminosity with accretion rate as the only relevant factor. It is now easy to see that if irradiation is very important in T~Pyx and has caused us to overestimate the accretion rate by a significant amount, then the non-detection of jet lines in the spectra of other CVs does become inconsistent with the presence of jets in these systems. Indeed, if $f_{\\dot{M}}\\gtappeq 4$, then even the previously derived lower limit of 0.06~\\AA~on the jet line EWs in (suitably selected) CVs becomes as large as 1~\\AA~and thus comparable to the strength of the same features in T~Pyx. In practical terms, this means that studies of T~Pyx aimed at deriving $T_{WD}$ (or, more precisely, $L_{WD}/L_{acc}$) for this system may provide yet another way to falsify our working hypothesis observationally in the future. \\subsection{The scaling relation applied to T~Pyx and the SSSs} \\label{BLA3} Given that the jets in T~Pyx and the SSSs are presumably of the same type, it is natural to try and use these systems to check our scaling relation for the jet line EWs. Unfortunately, the accretion rates of the relevant SSSs are only poorly constrained and, in addition, disk irradiation by the hot WD is likely to be very strong in the SSSs. As a consequence, a rigorous test of Relation~(\\ref{final2}) via this route is not possible. However, we will nevertheless proceed to apply our scaling relation to T~Pyx and the SSS RX~J0513-069, partly to illustrate these problems, and partly to perform at least a rough consistency check. In their study of RX~J0513-069, Southwell et al. (1996) state that $\\dot{M}_{acc} \\sim 10^{-5}$~M$_{\\sun}$~yr$^{-1}$ is required if the optical luminosity of this system is to be ascribed entirely to a standard accretion disk. An accretion rate this high is of the order of the Eddington value, and Southwell et al. (1996) therefore conclude that it is almost certainly an overestimate. They argue that irradiation of the disk and secondary star, as well as (perhaps) direct light from the hot WD are likely to contribute significantly to the optical light. Consequently, they prefer a lower value of about $10^{-6}$~M$_{\\sun}$~yr$^{-1}$ for the accretion rate. To make progress in the face of this uncertainty, we will adopt the higher value to start with and then check {\\em a posteriori} what value this implies for the correction factor $f_{\\dot{M}}^2$. Regarding RX~J0513-069's other relevant parameters, Southwell et al. (1996) give values of $P_{orb} \\simeq 18$~hrs for the orbital period, and, based on the mass function of the system, $i \\simeq 10^o$ for the inclination. Adopting these parameters for RX~J0513-069, and using the same parameters as above for T~Pyx, we would predict a best-bet ratio for the EWs of the jet lines in these two systems of about 50 (in favor of the SSS). Now, Southwell et al. (1996) measure the equivalent widths of the blue and red H$\\alpha$~jet satellite lines in RX~J0513-069 to be EW(SSS,blue) $\\simeq 1.6$~\\AA~and EW(SSS,red) $\\simeq 2.6$~\\AA, respectively. Thus the actual ratio of the jet line EWs in RX~J0513-069 and T~Pyx is only about 2. If we interpret this as a result of disk irradiation in the SSS, then the correction factor $f_{\\dot{M}}^2 \\simeq 25$ and $f_{\\dot{M}} \\simeq 5$. Consequently, we would predict the true accretion rate in RX~J0513-069 to be about $\\dot{M}_{acc} \\sim 2 \\times 10^{-6}$~M$_{\\sun}$~yr$^{-1}$, which is in line with the value of $10^{-6}$~M$_{\\sun}$~yr$^{-1}$ preferred by Southwell et al. (1996). We do not attach too much weight to this apparent consistency, because there are large observational uncertainties associated with the ratios constructed from two of the relevant parameters (accretion rate and inclination). Moreover, the accretion rate and orbital period ratios are so large for these systems that the theoretical uncertainties expressed by the ``errors'' in Relation~(\\ref{final2}) also become rather large. It is finally interesting to consider briefly the implications of adopting the complement of our working hypothesis. Specifically, we may ask whether a consistent physical picture capable of accounting for the relative strengths of the jet lines in T~Pyx and RX~J0513-069 can also be found if we assume that a hot, central object is in fact present in both systems and is crucial for driving the observed jets. To answer this question, we take $\\dot{M}_{jet} \\propto L_{WD}$ and assume the extreme case of $L_{WD} >> L_{acc}$. The disk is then still quite likely to dominate the optical flux (c.f. the numerical estimates for the direct WD contribution given previously), but its local temperature distribution will be dominated by irradiation, not dissipation (see Section~\\ref{BLA2}). Since the disk will be extremely hot in this case, we may further assume that the optical waveband lies on the Rayleigh-Jeans tail of the disk spectrum now, i.e. $F_{opt} \\propto L_{disk}^{1/4} \\propto L_{WD}^{1/4}$ (the latter holds since $L_{disk}$ is now dominated by $L_{WD}$). We can then replace the dependence on $\\dot{M}_{acc}$ in Relation~(\\ref{final2}) with one on $L_{WD}$, giving $EW(line) \\propto L_{WD}^{7/4} \\propto T_{WD}^7$. Adopting again an EW ratio of 2 for RX~J0513-069 and T~Pyx, we find that $T_{WD}(SSS) \\simeq 2~T_{WD}(T~Pyx)$ in this simplistic picture, if the remaining parameter dependences in Relation~(\\ref{final2}) are assumed to stay unchanged. This reasonable looking result should of course not be taken too seriously. However, the moral of this simple calculation is that it is certainly possible to account for the jet line EW differences between T~Pyx and RX~J0513-069 in the context of a model in which the presence of an energy source at disk center {\\em is} a crucial ingredient in driving the observed jets. This prompts us to stress again that our analysis in this paper has only shown that the presence of jets in CVs should not be ruled out simply because no jet lines have so far been detected in the optical spectra of these systems. We have by no means demonstrated that jets are actually present, or are even likely to be present, in ordinary CVs. {\\bf Note added:} After this paper was accepted for publication, we received a draft of a work by Margon \\& Deutsch, in which it is argued that the satellite lines seen in T~Pyx are in fact due to [N~{\\sc ii}]~$\\lambda\\lambda$6548,6584 and are formed in the complex velocity field of T~Pyx's nova shell(s). While the analytic scaling relation we derived in Section~2 retains its validity (and, we believe, usefulness) if this interpretation turns out to be correct, the same is not true for the prediction we made for the jet line EWs in ordinary CVs (since this is based on the assumption that T~Pyx's satellite lines are jet features). The best we can do in this case is to derive a new prediction by scaling down directly from one of the SSSs to CVs. To do this, we use Southwell~et al.'s (1997) inclination, orbital period and accretion rate estimates for RX~J0513-069 ($i\\simeq 10^o$; $P_{orb} \\simeq 18$~hrs; $\\dot{M}_{acc}(apparent)\\sim 10^{-5}$~M$_{\\sun}$~yr$^{-1}$ with $f_{\\dot{M}}=10$) and, as before, parameters appropriate to an optimally selected, ``ordinary'' CV ($i\\simeq 70^o$; $P_{orb} \\simeq 1.8$~hrs; $\\dot{M}_{acc} \\sim 10^{-8}$~M$_{\\sun}$~yr$^{-1}$). Ignoring limb-darkening ($\\eta(i)=1$), we obtain a new prediction of $EW(CV) \\sim 0.3$~\\AA. Even though the uncertainties on this number are substantial and hard to quantify (see Section~2.3), this estimate still suggests it would be premature to rule out the presence of jets in CVs completely at this stage. \\footnote{Note that if jets are present in CVs but jet lines are not seen in T~Pyx, the {\\em absence} of the latter would have to be attributed to one or both of the following: (i) T~Pyx's inclination is much lower than $70^o$; (ii) irradiation is increasing the brightness of the accretion disk in T~Pyx substantially.} We therefore suggest that an optical survey of suitably selected CVs to search for jet lines is called for, regardless of the nature of the satellite lines in T~Pyx." }, "9803/astro-ph9803071_arXiv.txt": { "abstract": " ", "introduction": " ", "conclusions": "These two measurements of D/H in QSO absorption systems are the best and most robust measures to date. Deuterium has been identified and analyzed in a number of other QSO absorption systems\\cite{dhother} We have found another two systems which place a strong upper limit on D/H at D/H $< 10^{-4}$. Combined with the two measurements described above, the four independent systems support a low primordial abundance of deuterium, and together give D/H = 3.4 $\\pm \\, 0.3 \\times 10^{-5}$. If this represents the primordial value, nucleosynthesis calculations from standard BBN models with three light neutrinos give $\\eta = 5.1 \\pm \\, 0.3 \\times 10^{-10}$ and $\\Omega_b\\,h_{100}^2 = 0.019 \\pm \\, 0.001$. The constraints from D/H can be utilized to constrain cosmological models, quantify dark matter both in unobserved baryons and non-baryons, specify the zero point for models of deuterium evolution\\cite{chemevol}, test directly the predictions of standard BBN by comparing with other light element abundances\\cite{hat}, and limit the amount of small scale entropy fluctuations in the early universe\\cite{jed}." }, "9803/astro-ph9803247_arXiv.txt": { "abstract": "We have obtained spectra of the Galactic center at energies 400--600 keV from high-resolution data acquired by the TGRS Ge spectrometer on board the {\\em WIND\\/} mission during 1995--1997. The data were obtained using an on-board occulter, and are relatively free from systematics and backgrounds. Analysis of the spectra reveals a well-resolved electron-positron annihilation line at 511 keV and the associated continuum due to annihilation via positronium formation. Measurements of the line width and the continuum-to-line ratio allow some constraints to be placed on the interstellar sites where annihilation occurs. ", "introduction": "The line at 511 keV from the annihilation of electrons and positrons in the region of the Galactic center (GC) is the best-studied line in $\\gamma$-ray astronomy. Over 20 years of observations (reviewed by Tueller 1993) have established that there is an extensive diffuse line source of total intensity $\\sim 2 \\times 10^{-3}$ photon cm$^{-2}$ s$^{-1}$. This source has recently been mapped in considerable detail by OSSE on board the {\\em Compton Observatory\\/} (Purcell et al. 1997), which has revealed a third spatial component in addition to the well-known Galactic disk and bulge components. This new component is extended and is centered at $l = -2^{\\circ}$, $b = +9^{\\circ}$, well above the Galactic plane. It is unclear whether there are any point sources superimposed on this diffuse distribution; recent results do not show any variability in the flux. The line is known to be narrow and centered at 511 keV (Leventhal, MacCallum, \\& Stang 1978). The annihilation spectrum also includes a lower-energy continuum arising from $3 \\gamma$ annihilation via the formation of positronium (Ps). In principle, spectral lines contain much information about the physical conditions in the line formation region. The next step in the study of the 511 keV line will be to extract the information contained in the line profile and in the ratio of line to Ps continuum amplitudes. The key requirement is for sensitive long-term measurements with fine spectral resolution. The measurements described above were mostly made with low-resolution scintillator detectors. In this paper, we describe observations made over more than 2 years with the high-resolution Ge spectrometer TGRS on board the {\\em WIND\\/} spacecraft. ", "conclusions": "The results of our measurements of the annihilation spectrum are given in Table 1. These results supersede the preliminary measurement made by Teegarden et al. (1995), which reported a 511 keV line flux $1.64 \\times 10^{-3}$ photon cm$^{-2}$ s$^{-1}$. Two features of the earlier analysis contributed to this overestimate. First, the line flux was obtained from the count rate in the narrow 506--516 keV band, without fitting the shape of the spectrum, and is thus overestimated by including the other spectrum components (Fig. 1). Second, the present analysis uses an improved model of the instrument spectral response, instead of simply dividing by photopeak effective area as was done in the earlier work. \\subsection{Comparison with OSSE results} We calculated the flux, dimension and centroid of the OSSE model of the GC 511 keV emission (Purcell et al. 1997) as folded through the TGRS occulted response (Table 1). Since the occulter passes close to the centers of two of the three spatial components of the model (exactly crossing the GC, and $2.8^{\\circ}$ from the high-latitude feature) a test of these model features becomes possible in principle. The measurements of the flux and centroid are in good agreement; the offset of the TGRS centroid measurement from the GC is in the same direction as the offset of the OSSE centroid due to the new high-latitude feature, but is also compatible with the GC. The spatial extension found by TGRS slightly exceeds that found by OSSE, but to draw any conclusion from this would be premature since improved modeling of the occultation response is required. Our results agree with OSSE in finding a lack of variability on 90 d time-scales (Fig. 3). \\subsection{Source physics: Line width} Our measurements of the total line width $\\sigma_{tot}$, and of the background line widths, are shown in Fig. 4. The interpolated instrument intrinsic width $\\sigma_{inst}$ is also shown. It is clear that $\\sigma_{tot}$ at all times exceeds $\\sigma_{inst}$; this is necessary if the cosmic line width is to be obtained from $\\sigma_{tot}^{2} = \\sigma_{inst}^{2} + \\sigma_{gal}^{2}$. The result (Table 1) is somewhat narrower than the average of four balloon measurements by the GRIS Ge detector (Leventhal et al. 1993), but the difference is not very significant. The line width reflects the convolved widths of components due to different annihilation mechanisms predominating in different phases of the ISM. These mechanisms were treated by Guessoum, Ramaty \\& Lingenfelter (1991). They may be divided into two classes. Firstly, annihilation by charge-exchange in flight produces a broad line (FWHM 6.4 keV), and is predominant in cold molecular clouds. The second class contains all other processes, which produce lines narrower than the instrument resolution. We can therefore hope to test two alternative suggestions by Guessoum et al. --- annihilation occurring uniformly in all phases of the ISM, and otherwise-uniform annihilation excluding cold clouds. We therefore repeated our analysis under the assumption that the width $\\sigma_{gal}$ had two components, a broad component of width 6.4 keV, and a narrow unresolved component. Instead of line width, we now have the amplitude of the 6.4-keV broad component as a fitted parameter. The spectra were fitted equally well by this model; though there were no significant improvements in the $\\chi^2$ values, we hope this procedure yields physical insight into the meaning of $\\sigma_{gal}$. Assuming the presence of a 6.4-keV broad line component, we found that $11$\\%$\\pm 9$\\% of the total line intensity was due to this broad line. This is much closer to the prediction when positrons are excluded from molecular cloud cores (in which case the broad line contributes only 11\\%: Guessoum et al. 1991) than to the maximum predicted broad-line contribution of 59\\% when positrons penetrate all phases of the ISM equally. \\subsection{Source physics: Positronium fraction} The fraction $f$ of positrons which annihilate through the formation of Ps can be written $f = 2/[2.25(I_{511}/I_{Ps})+1.5]$, where $I_{511}$ and $I_{Ps}$ are the line and Ps continuum intensities (Brown \\& Leventhal 1989); our result from Table 1 is $f = 0.94 \\pm 0.04$.\\footnote{ The uncertainty does not include those systematic errors in $I_{511}$ and $I_{Ps}$ in Table 1 which are positively correlated.} This is in good agreement with the most recent OSSE result $f = 0.97 \\pm 0.03$ (Kinzer et al. 1996). However, predicted values from annihilation in most of the phases of the ISM cluster in the range $f \\sim 0.9$--1.0, so small discrepancies in measured $f$ may be important. Our result falls roughly in the middle of this range, and is consistent with annihilation in cold molecular clouds ($f = 0.9$: Brown, Leventhal \\& Mills 1986), the warm neutral or ionized ISM ($f = 0.9$--0.95: Bussard, Ramaty \\& Drachman 1979), or any combination of these (Guessoum et al. 1991). It is not compatible with annihilation in the hot phase, nor with any scenario in which grains are important sites of annihilation. These two statements are in fact equivalent, since in the hot phase grains become the most important location for annihilation in the absence of H atoms. The corresponding value of $f$ is expected to be very low ($\\le 0.5$: Guessoum et al. 1991). \\subsection{Summary} We have measured the 511 keV line from the GC and also the Ps continuum associated with it during 1995--1997. Our values for the intensities of these features agree with the most recent OSSE measurements. Our preliminary results for the spatial distribution of the line are consistent with the OSSE mapping, but require further analysis of the instrument response. The 511 keV line is resolved, and, if a specific model for its width is assumed (an underlying broad component from annihilation through charge-exchange in flight), then our result favors a scenario in which annihilation in cold molecular clouds is suppressed. Our measurement of the Ps fraction $f$ from the Ps continuum is consistent with this, and suggests further that annihilation in the hot phase of the ISM is of minor importance." }, "9803/astro-ph9803137_arXiv.txt": { "abstract": "We report the discovery of \\lya\\ emission from a galaxy at $z=5.34$, the first object at $z>5$ with a spectroscopically confirmed redshift. The faint continuum emission (${\\rm m_{AB}(8000{\\rm \\AA})\\approx 27}$), relatively small rest-frame equivalent width of the emission line ($W_{\\rm Ly\\alpha}^{rest}\\approx 95$\\AA), and limits on the \\ion{N}{5}/\\lya\\ ratio suggest that this is a star--forming galaxy and not an AGN. The star--formation rates implied by the UV continuum emission and the \\lya\\ emission are (in the absence of dust extinction) fairly modest ($\\sim 6~h_{50}^{-2}\\ \\Msun~yr^{-1}$ for \\qnot=0.5). The continuum luminosity is similar to that of sub-$L^*_{1500}$ star--forming galaxies at $z\\sim3$, and the width of the \\lya\\ line yields an upper limit to the mass of $< 2.6\\times 10^{10}\\Msun$. The strong emission line detected in this low-luminosity galaxy provides hope for the discovery of higher luminosity primeval galaxies at redshifts $z>5$. ", "introduction": " ", "conclusions": "" }, "9803/astro-ph9803301_arXiv.txt": { "abstract": "Using the Hubble Space Telescope and WFPC2 we have imaged the central 20pc of the giant H~II region 30 Doradus nebula in three different emission lines. The images allow us to study the nebula with a physical resolution that is within a factor of two of that typical of ground based observations of Galactic H~II regions. We present a gallery of interesting objects within the region studied. These include a tube blown by the wind of a high velocity star and a discrete H~II region around an isolated B star. This small isolated H~II region appears to be in the midst of the champagne flow phase of its evolution. Most of the emission within 30 Dor is confined to a thin zone located between the hot interior of the nebula and surrounding dense molecular material. This zone appears to be directly analogous to the photoionized photoevaporative flows that dominate emission from small, nearby H~II regions. For example, a column of material protruding from the cavity wall to the south of the main cluster is found to be a direct analog to elephant trunks in M16. Surface brightness profiles across this structure are very similar to surface brightness profiles taken at ground based resolution across the head of the largest column in M16. The dynamical effects of the photoevaporative flow can be seen as well. An arcuate feature located above this column and a similar feature surrounding a second nearby column are interpreted as shocks where the photoevaporative flow stagnates against the high temperature gas that fills the majority of the nebula. The ram pressure in the photoevaporative flow, derived from thermal pressure at the surface of the column, is found to balance with the pressure in the interior of the nebula derived from previous x-ray observations. By analogy with the comparison of ground and HST images of M16 we infer that the same sharply stratified structure seen in HST images of M16 almost certainly underlies the observed structure in 30 Dor. 30 Doradus is a crucial case because it allows us to bridge the gap between nearby H~II regions and the giant H~II regions seen in distant galaxies. The real significance of this result is that it demonstrates that the physical understanding gained from detailed study of photoevaporative interfaces in nearby H~II regions can be applied directly to interpretation of giant H~II regions. Stated another way, interpretation of observations of giant H~II regions must account for the fact that this emission arises not from expansive volumes of ionized gas, but instead from highly localized and extremely sharply stratified physical structures. ", "introduction": "30 Doradus is a giant ionized complex in the Large Magellanic Cloud (LMC), located at a distance of 51.3 kpc (eg. Panagia et al 1991). The nebula is centered on a dense cluster of newly formed stars, the most dense component of which is called R136. The nebula itself is more than 180 parsecs across, qualifying it as a smaller member of the elite class of nebulae termed Giant Extragalactic H~II Regions (GEHR's). If 30 Doradus was placed at the distance of the Orion Nebula from the Earth, it would appear to be more than 20 degrees across, and would fill more than 4\\% of the night sky. The central cluster is very dense and is comprised of several hundred OB stars with a small number of W-R stars (Hunter et al 1995b). The integrated ultraviolet flux from this cluster is intense: more than fifty times that being produced in the center of the Orion Nebula (Campbell et al 1992). Radiation from the cluster combined with strong stellar winds from the most massive stars in the cluster has eroded a large cavity in the nearby molecular complex, producing the nebula we see today. Hunter et al 1995b showed that the majority of the stars in the cluster were formed in a single star formation event more than 2-3 million years ago. The census performed yielded a ``head count\" of more than 3000 stars with more than 300 OB stars capable of producing the intense UV radiation and stong stellar winds responsible for forming and shaping the H~II regions we observe in galaxies. The level of star formation exhibited by the 30 Doradus region and the neighboring LMC complex are the closest example of starburst-like star formation. As such we are getting a unique view of the star formation environment in the middle of an ongoing starburst. The average reddening along the line of sight to the Large Magellanic Cloud and 30 Doradus is very low (Panagia et al 1991). However, within several H~II complexes in the LMC comparison between optical and radio measurements suggest a large variation in the local reddening. Kennicutt \\& Hodge 1986 found a variation in these estimates between 0 and 1 magnitude in A$_{V}$. Hunter et al 1995b also found substantial variation in the reddening across the face of the 30 Doradus nebula, and derived a mean estimate for this reddening of 1.4 magnitudes in A$_{V}$ at 555 nm, and 0.8 magnitudes at 814 nm. For the purposes of this paper we will adopt an extinction of 1 magnitude in A$_{V}$ for the emission lines we observe. The 30 Doradus nebula plays a key role in our understanding of H~II regions. Nearby regions are close enough for the physical processes at work within the nebula to be studied in detail. The work by Hester et al 1996 (hereafter H96) on M16, for example, shows that emission within the nebula arises predominantly within a narrow region at the interface between the H~II region and the molecular cloud. They follow Hester 1991 in describing this thin region as a photoionized photoevaporative flow. However, an H~II region like M16 is tiny in comparison with giant H~II regions, and no giant H~II regions are close enough to allow the stratified ionization structure of the photoevaporative flow to be studied directly. 30 Doradus alone offers an opportunity to bootstrap the physical understanding of small nearby H~II regions into the context of the giant regions seen in distant galaxies. In this paper we present Hubble Space Telescope images of the ionization structure we observe around the central cluster R136. The wealth of spatial information contained in these pictures is daunting to consider, but we attempt to summarize the most telling points by selecting and presenting several examples of distinct structures around the field of view that provide insight into how the interface with the local gas and dust is evolving. In \\S 2 we discuss the observations themselves and the general structure of the nebula, as well as presenting full-field mosaics of the data. In \\S 3 we discuss the conditions apparent in the ionized hydrogen along the walls of the H~II region cavity, as well as comparing the ionization structure we observe with models we derive from the H$\\alpha$ surface brightness. ", "conclusions": "We have presented high resolution narrow-band imagery of the 30 Doradus nebula. There are many interesting localized structures within the nebula, a number of which appear to be associated with winds and UV from stars that are not part of the main 30 Doradus cluster. However the majority of the emission from the nebula is due to photoionization by the flux from the central cluster. This emission is largely concentrated in thin regions located at the interface between dense molecular material and the shock-heated interior of 30 Dor. At the resolution of the {\\it HST} data we find that the structure in 30 Doradus is remarkably similar to what is seen in ground-based observations of nearby H~II regions. This similarity is not surprising given that despite an overall difference in scale, locally the physical conditions in 30 Doradus are not much different than those found in smaller H~II regions. We demonstrate this point above by focussing on one particular region in 30 Doradus and showing that it is a very direct analog of the Galactic H~II region M~16. Taking this same argument a step further we are lead to the conclusion that underlying the observed structure in M~16 is the same sort of extremely localized and sharply stratified structure seen in the {\\it HST} images of M~16. Thus, even though the 30 Doradus nebula spans hundreds of parsecs, the emission from this giant H~II region arises largely in the same sorts of sharply stratified photoionized photoevaporative flows seen in nearby H~II regions. The 30 Doradus nebula is a crucial case. The fact that at {\\it HST} resolution 30 Doradus is so similar to ground based images of nearby H~II regions has allowed us to bootstrap our physical understanding based on detailed study of nearby regions into the physical context of a giant H~II region surrounding a massive young cluster. Similarly, preliminary analysis of {\\it HST} images of more distant giant H~II regions suggests that they compare favorably with 30 Dor when that nebula is viewed at the same physical resolution. This indicates that conditions in 30 Doradus are probably typical of those found in these distant H~II regions as well. Bootstrapping first from nearby H~II regions to 30 Doradus in this paper, and we anticipate from 30 Doradus to more distant regions in later work, we are approaching the conclusion that the emission from giant H~II regions megaparsecs distant is determined by the physics of photoevaporative flows in which relevant physical scales can be as small as 100 AU or less. The significance of this work lies in the conclusion that the detailed study of nearby, well-resolved H~II regions is directly applicable to distant giant H~II regions in much the same way that an understanding of radiative shocks that is tested in nearby supernova remnants can be applied in a variety of contexts in which the shock itself is not resolved. Viewed from a different perspective, interpretation of observations of distant giant H~II regions must take into account the fact that much of this emission arises not in vast expanses of ionized or even clumpy gas, but instead in well defined and highly stratified photoevaporative flows localized to the surfaces of molecular clouds." }, "9803/astro-ph9803315_arXiv.txt": { "abstract": "The process of molecule formation in the primordial gas is considered in the framework of Friedmann cosmological models from redshift $z=10^4$ to $z=0$. First, a comprehensive analysis of 87 gas phase reaction rates (both collisional and radiative) relevant in the physical environment of the expanding universe is presented and critically discussed. On this basis, calculations are carried out of the abundance of 21 molecular species as function of redshift, for different values of the cosmological parameters $\\Omega_0$, $\\eta$ and $H_0$, evaluating consistently the molecular heating and cooling due to H$_2$, HD and LiH molecules. One of the major improvements of this work is the use of a better treatment of H recombination that leads to a reduction of a factor 2--3 in the abundance of electrons and H$^+$ at freeze-out, with respect to previous studies. The lower residual ionization has a negative effect on the chemistry of the primordial gas in which electrons and protons act as catalysts in the formation of the first molecules. We find that in the standard model ($h=0.67$, $\\eta_{10}=4.5$, $\\Omega_0=1$ and [D/H] $=4.3\\times 10^{-5}$), the residual fractional ionization at $z=1$ is $[{\\rm e/H}]=3.02\\times 10^{-4}$, and the main molecular species fractional abundances $[{\\rm H}_2/{\\rm H}]=1.1\\times 10^{-6}$, $[{\\rm HD/H}_2]=1.1\\times 10^{-3}$, $[{\\rm HeH}^+/{\\rm H}]=6.2\\times 10^{-13}$, $[{\\rm LiH}^+/{\\rm H}]=9.4\\times 10^{-18}$ and $[{\\rm LiH/LiH}^+]=7.6\\times 10^{-3}$. We devise a reduced chemical network that reproduces with excellent accuracy the numerical results of the complete model and allows to follow the chemical compositions and the thermal properties of a primordial gas in the presence of an external radiation field. Finally, we provide accurate cooling functions of H$_2$, HD and LiH in a wide range of density and temperature that can be conveniently used in a variety of cosmological applications. ", "introduction": "The study of molecule formation in the post-recombination epoch has grown considerably in recent years. Saslaw \\& Zipoy (1967) and Peebles \\& Dicke (1968) were the first to realize the importance of gas phase reactions for the formation of the simplest molecule, H$_2$. They showed that trace amounts of molecular hydrogen, of order 10$^{-6}$--10$^{-5}$, could indeed form via the intermediaries species H$_2^+$ and H$^-$ once the radiation field no longer contained a high density of photons with energies above the threshold of dissociation (2.64 and 0.75 eV, respectively). The presence of even a trace abundance of H$_2$ is of direct relevance for the cooling properties of the primordial gas which, in its absence, would be an extremely poor radiator: cooling by Ly-$\\alpha$ photons is in fact ineffective at temperatures $\\la 8000$~K, well above the matter and radiation temperature in the post-recombination era. Since the evolution of primordial density fluctuations is controlled by the ability of the gas to cool down to low temperatures, it is very important to obtain a firm picture of the chemistry of the dust-free gas mixture, not limited to the formation of H$_2$, but also to other molecules of potential interest. In this regard, Lepp \\& Shull (1984) and Puy et al. (1993) have computed the abundances of H$_2$, HD and LiH as a function of redshift for various cosmological models. Although the final abundances of H$_2$ and HD agree in the two calculations, their evolution with redshift is markedly different, since the epoch of formation varies by a factor of $\\sim 2$. Also, the LiH abundance shows a large discrepancy of about two orders of magnitude. More recently, Palla et al. (1995) have analyzed the effects on the chemistry of the pregalactic gas of a high primordial D abundance in the light of the controversial results obtained towards high redshift quasars (see e.g. Tytler \\& Burles 1997). They found that the abundance of H$_2$ is rather insensitive to variations in the cosmological parameters implied by a factor of $\\sim 10$ enhancement of primordial [D/H], while HD and LiH abundances vary by larger amounts. However, the abundance of LiH, obtained with simple estimates of the radiative association rate, was largely overestimated. Because of the potential relevance of the interaction of LiH molecules with the cosmic background radiation (CBR) (Maoli et al. 1994), a proper treatment of the lithium chemistry was necessary. Dalgarno et al. (1996) and Gianturco \\& Gori Giorgi (1996a,b) provided accurate quantum-mechanical calculations of the main reaction rates. The chemistry of lithium in the early universe has been then studied by Stancil et al. (1996) and Bougleux \\& Galli (1997). Finally, useful reviews of the chemistry of the early universe can be found in Dalgarno \\& Lepp (1987), Black (1991), Shapiro (1992), and Abel et al. (1997). The latter two studies, in particular, focus on the nonequilibrium H$_2$ chemistry in radiative shocks which is thought to be of primary importance during the gravitational collapse of density fluctuations (see also Anninos et al. 1997). In spite of such a wealth of specific studies, a comprehensive analysis of the subject and a critical discussion of the reaction paths and rates are still lacking. To overcome this limitation, in this paper we present a complete treatment of the evolution of {\\em all} the molecular and atomic species formed in the uniform pregalactic medium at high redshifts ($z<10^4$). The structure of the paper is as follows: in Sect.~2 we describe the H, D, He, and Li chemistry, with a critical discussion of the most important rates; the evolutionary models are presented in Sect.~3, and the results for the standard model and the dependence on the cosmological parameters are given in Sect.~4; Sect.~5 introduces a minimal model which highlights the dominant reactions for the formation of H$_2$, HD, HeH$^+$, LiH and LiH$^+$; a comparison with the results of previous studies is given in Sect.~6, and the conclusions are summarized in Sect.~7. Also, the Appendix provides the collisional excitation coefficients for HD and H$_2$ and cooling function of H$_2$, HD and LiH which are needed for the computation of the thermal evolution of the primordial gas. ", "conclusions": "The main results of the present study can be summarised as follows: 1) We have followed the chemical evolution of the primordial gas after recombination by computing the abundances of 21 species, 12 atomic and 9 molecular, by using a complete set of reaction rates for collisional and radiative processes. The rates which are critical for a correct estimate of the final molecular abundances have been analysed and compared in detail. 2) One of the major improvements of this work is the use of a better treatment of H recombination that leads to a reduction of a factor 2--3 in the abundance of electrons and H$^+$ at freeze-out, with respect to previous studies. The lower residual ionization has a negative effect on the chemistry of the primordial gas in which electrons and protons act as catalysts in the formation of the first molecules. 3) In the standard model ($h=0.67$, $\\eta_{10}=4.5$, $\\Omega_0=1$ and [D/H] $=4.3\\times 10^{-5}$), the residual fractional ionization at $z=1$ is $[{\\rm e/H}]=3.02\\times 10^{-4}$, and the main molecular species have fractional abundances $[{\\rm H}_2/{\\rm H}]=1.1\\times 10^{-6}$, $[{\\rm HD/H}_2]=1.1\\times 10^{-3}$, $[{\\rm HeH}^+/{\\rm H}]=6.2\\times 10^{-13}$, $[{\\rm LiH}^+/{\\rm H}]=9.4\\times 10^{-18}$ and $[{\\rm LiH/LiH}^+]=7.6\\times 10^{-3}$. 4) As for molecular hydrogen, its final abundance does not depend on the model parameters, making this molecule a poor diagnostic of cosmological scenarios. The largest uncertainty resides in the accurate knowlwedge of the photodissociation rate of H$_2^+$. A detailed treatment of the reaction kinetics of this reaction would be required. 5) We have presented a minimal model consisting of 11 reactions for H$_2$, 6 for HD, 3 for HeH$^+$ and 14 for LiH which reproduces with excellent accuracy the results of the full chemical network, regardless of the choice of the cosmological parameters. 6) Finally, we have computed accurate expressions for the cooling functions of H$_2$, HD and LiH in a wide range of density and temperature that can be conveniently used in a variety of comological applications." }, "9803/astro-ph9803123_arXiv.txt": { "abstract": "We have obtained WFPC2 images of 256 of the nearest (z$\\leq$0.035) Seyfert 1, Seyfert 2, and starburst galaxies. Our 500-second broadband (F606W) exposures reveal much fine-scale structure in the centers of these galaxies, including dust lanes and patches, bars, rings, wisps and filaments, and tidal features such as warps and tails. Most of this fine structure cannot be detected in ground based images. We have assigned qualitative classifications for these morphological features, a Hubble type for the inner region of each galaxy, and also measured quantitative information such as 0.18 and 0.92 arcsecond aperture magnitudes, position angles and ellipticities where possible. There is little direct evidence for unusually high rates of interaction in the Seyfert galaxies. Slightly less than 10\\% of all the galaxies show tidal features or multiple nuclei. The incidence of inner starburst rings is about 10\\% in both classes of Seyfert galaxies. In contrast, galaxies with H II region emission line spectra appear substantially more irregular and clumpy, because of their much higher rates of current star formation per unit of galactic mass. The presence of an unresolved central continuum source in our {\\it HST} images is a virtually perfect indicator of a Seyfert 1 nucleus as seen by ground-based spectroscopy. Fifty-two percent (52\\%) of these Seyfert 1 point sources are saturated in our images; we use their wings to estimate magnitudes ranging from 15.8 to 18.5. The converse is not universally true, however, as over a third of Seyferts with direct spectroscopic evidence for broad Balmer wings show no nuclear point source. These 34 resolved Seyfert 1's have fainter nonstellar nuclei, which appear to be more extinguished by dust absorption. Like the Seyfert 2's, they have central surface brightnesses consistent with those expected for the bulges of normal galaxies. The rates for the occurrences of bars in Seyfert 1's and 2's and non-Seyferts are the same. We found one significant morphological difference between the host galaxies of Seyfert 1 and Seyfert 2 nuclei. The Seyfert 2 galaxies are significantly more likely to show nuclear dust absorption, especially in lanes and patches which are irregular or reach close to the nucleus. A few simple tests show that the difference cannot be explained by different average redshifts or selection techniques. It is confirmed by our galaxy morphology classifications, which show that Seyfert 1 nuclei reside in earlier type galaxies than Seyfert 2 nuclei. If, as we believe, this is an intrinsic difference in host galaxy properties, it would undermine one of the postulates of the strong unification hypothesis for Seyfert galaxies, that they merely appear different due to the orientation of their central engine. The excess galactic dust we see in Seyfert 2's may cause substantial absorption which obscures their hypothesized broad-emission-line regions and central nonstellar continua. This galactic dust could produce much of the absorption in Seyfert 2 nuclei which had instead been attributed to a thick dusty accretion torus forming the outer part of the central engine. ", "introduction": "Several causal connections have been proposed between an active galactic nucleus (AGN) and the host galaxy in which it resides. The principal ways in which the latter could affect the former are through influencing a) the formation of a nonstellar central engine; b) its fueling; and c) obscuring it from our view, (which can alter the central engine's appearance even if it is not physically affected.) It is widely believed that active galactic nuclei (AGNs) are powered by non-spherical accretion onto massive black holes. This is partly because this model has the lowest fuel supply requirements: an AGN's luminosity is proportional to its mass accretion rate, which would be about 0.01 \\msun year$^{-1}$ for a bright Seyfert nucleus. It is not known how this rate of fuel supply can be brought from the host galaxy down to several thousand Schwarzschild radii (of order 10$^{17}$ cm for a ``typical\" Seyfert galaxy black hole mass of 10$^8$ \\msun (Malkan \\markcite{a60} 1983)) at which point viscous processes are supposed to drive the final accretion onto the black hole. One speculation is that a close interaction with another galaxy can distort the galactic potential and disturb the orbits of gas clouds sufficiently to carry a significant mass of fuel into the galaxy's center (Shlosman \\etal\\markcite{a1}\\markcite{a2}1989, 1990, Hernquist and Mihos \\markcite{a3}1996). More indirect scenarios are also possible, in which a tidal galaxy interaction stimulates a burst of star formation which in turn stimulates nonstellar nuclear activity. A further possibility is that special conditions in isolated galaxies may trigger the feeding of fuel to an active nucleus, such as a bar instability. (Schwartz \\markcite{a31}1981; Shlosman, Frank, \\& Begelman \\markcite{a2}1990; Mulchaey and Regan \\markcite{a32}1997) However, direct observational evidence that galaxy encounters stimulate the luminosity of an AGN has been ambiguous (Adams \\markcite{a4}1977, Petrosian \\markcite{a65}1983, Kennicutt and Keel \\markcite{a5}1984, Dahari \\markcite{a6}\\markcite{a7}1985a, 1985b, Bushouse \\markcite{a8}1986, Fuentes-Williams and Stocke \\markcite{a66}1988). One difficulty is that the most dramatic morphological indications of the encounter may have subsided by the time that the newly injected fuel reaches the nucleus. In any case, the weak correlation between galaxy interactions and Seyfert activity is stronger for type 2 Seyferts than for type 1's. Conversely, the presence of an AGN could alter the appearance of the central regions of its host galaxy, principally by its injection of substantial energy, both radiative and mechanical, over many millions of years. A further question is whether the particular type of active nucleus, Seyfert 1 or 2, is related to any property of the host galaxy. We have therefore used the superior spatial imaging resolution of the post-repair {\\it Hubble Space Telescope} to make a snapshot survey of nearby active galaxies to investigate the morphological implications of different theories on the formation and fueling of AGN. ", "conclusions": "Our large sample of high-resolution images of the centers of nearby Seyfert 1, 2 and HII galaxies has allowed us to search for statistical differences in their morphologies. The Seyfert galaxies do not, on average, resemble the HII galaxies. The latter have more irregularity and lumpiness associated with their high rates of current star formation. Conversely, none of the HII galaxies have the filaments or wisps which are sometimes seen in Seyfert 1 and 2 galaxies, and are evidently gas filaments photoionized by the active nucleus. Sixty-three percent (63\\%) of the galaxies classified as Seyfert 1 have an unresolved nucleus, 52\\% of which are saturated. Some (6\\%) have such dominant nuclei that they would appear as ``naked quasars\" if viewed at somewhat higher redshifts. The presence of an unresolved nucleus, particularly a saturated one, is anti-correlated with an intermediate spectroscopic classification (such as Seyfert 1.8 or 1.9) and is also anti-correlated with the Balmer decrement. This implies that those Seyfert 1's with weak nuclei in the PC2 images are extinguished and reddened by dust. The vast majority of the Seyfert 2 galaxies show no central point source. In fact, the only two of these that do (IRAS 1832-594 and IC 4870) are mis-classified galaxies. If all Seyfert 2's actually harbor point-like continuum sources like those in Seyfert 1's, they are at least an order of magnitude fainter on average. In those galaxies without any detectable central point source (37\\% of the Seyfert 1's; 98\\% of the Seyfert 2's, and 100\\% of the H II's), the central surface brightnesses are statistically similar to those observed in the bulges of normal galaxies. Seyfert 1's and 2's both show circumnuclear rings in about 10\\% of the galaxies. We identified strong inner bars as often in Seyfert 1 galaxies (27\\%) as in Seyfert 2 galaxies (22\\%). In some cases we see a strong assymetry of the dust absorption across the major axis, which allows us to infer which half of the disk is nearer to us: the side which more strongly absorbs the smooth light of the bulge behind it. The Seyfert 2 galaxies are more likely than Seyfert 1's to show irregular or disturbed dust absorption in their centers as well as galactic dust lanes which pass very near their nuclei. They also, on average, tend to have latter morphological types than the Seyfert 1's. This difference remains in Seyfert 1 and 2 subsamples matched for redshift, [OIII] and radio luminosities. It also holds true when we restrict our consideration to sub-samples of the data which are less biased by selection effects. Thus it appears that the host galaxies of Seyfert 1 and 2 nuclei are {\\it not} intrinsically identical. A galaxy with more nuclear dust and in particular more irregularly distributed dust is more likely to harbor a Seyfert 2 nucleus. This indicates that the higher dust-covering fractions in Seyfert 2's are the reason for their spectroscopic classification: their compact Seyfert 1 nucleus may have been obscured by galactic dust. This statistical result contradicts the simplest and most popular version of the unified scheme for Seyfert galaxies. We suggest that the obscuration which converts an intrinsic Seyfert 1 nucleus into an apparent Seyfert 2 often occurs in the host galaxy hundreds of parsecs from the nucleus. If so, this obscuration need have no relation to a hypothetical fat dust torus surrounding the equator of the central engine. Also then the orientation of the central engine with respect to our line-of-sight does {\\it not} determine whether an active nucleus will appear to us as a Seyfert 1 or as a Seyfert 2." }, "9803/astro-ph9803065_arXiv.txt": { "abstract": "Wind-blown bubbles, from those around massive O and Wolf-Rayet stars, to superbubbles around OB associations and galactic winds in starburst galaxies, have a dominant role in determining the structure of the Interstellar Medium. X-ray observations of these bubbles are particularly important as most of their volume is taken up with hot gas, $10^{5} \\ltsimm T (\\K) \\ltsimm 10^{8}$. However, it is difficult to compare these X-ray observations, usually analysed in terms of single or two temperature spectral model fits, with theoretical models, as real bubbles do not have such simple temperature distributions. Spectral fits, and the properties inferred from them, will depend in a complex way on the true temperature distribution and the characteristics and limitations of the X-ray observatory used. In this introduction to a series of papers detailing the {\\em observable} X-ray properties of wind-blown bubbles, we describe our method with which we aim to solve this problem, analysing a simulation of a wind-blown bubble around a massive star. Our model is of a wind of constant mass and energy injection rate, blowing into a uniform ISM, from which we calculate X-ray spectra as would be seen by the {\\it ROSAT} PSPC. Analysing these spectra in the same way as a real observation would be, we compare the properties of the bubble as would be inferred from the {\\it ROSAT} data with the true properties of the bubble in the simulation. We find standard spectral models yield inferred properties that deviate significantly from the true properties, even though the spectral fits are statistically acceptable, and give no indication that they do not represent to true spectral distribution. For example, single temperature spectral fits give best fit metal abundances only 4\\% of the true value. A cool bubble has best fit temperatures significantly higher than a bubble twice as hot. These results suggest that in any case where the true source spectrum does not come from a simple single or two temperature distribution the ``observed'' properties cannot naively be used to infer the true properties. In this situation, to compare X-ray observations with theory it is necessary to calculate the {\\em observable} X-ray properties of the model. ", "introduction": "Bubbles blown in the Interstellar Medium (ISM) by massive stars are a common astrophysical phenomenon. X-ray observations can provide information regarding the density, metal abundance, temperature, ionisation state and physical structure in the hot bubbles surrounding Wolf-Rayet (WR) and O stars (Wrigge, Wendker \\& Wisotski 1994), Luminous Blue Variables (LBV's) such as $\\eta$ Carinae (Corcoran \\etal 1995) and planetary nebulae (PN) (Kreysing \\etal 1992; Leahy, Zhang \\& Kwok 1994; Arnaud, Borkowski \\& Harrington 1996; Leahy \\etal 1996). On the larger scale, superbubbles are created by the sum of the winds and SN within OB associations (Belloni \\& Mereghetti 1994) and giant star forming regions in young starburst galaxies. Superbubbles within starburst galaxies such as M82 eventually break out the galaxy to form spectacular galactic winds (Watson, Stanger \\& Griffiths 1984; Heckman, Armus \\& Miley 1987). In many cases the X-ray emission probes different regions of the object in question to that revealed by optical observations, increasing the importance of the X-ray data. Analytic solutions to the development and structure of astrophysical bubbles must rely on simplifying assumptions, and increasingly attention has turned to the use of numerical hydrodynamics. These simulations have been enlightening with respect to the nonlinear processes occurring, with some degree of quantitative agreement with observation, but generally lack predictive power. This is partially due to the difficulty in comparing them with observations, in particular X-ray observations. The problem is that X-ray observations are usually analysed by fitting a single or two temperature spectral model to the observed spectra (see for example the references above), and the best-fit results are used to infer the physical properties of the object. However, for wind-blown bubbles such as those mentioned above, the true situation is more complex, and the results of the spectral fits may be influenced by, for example: projection of different physical regions along the line of sight; the presence of a wide range of temperatures; interstellar absorption; unknown or non-standard elemental abundances; non-ionisation equilibrium conditions; low numbers of observed photons and the limitations of the current X-ray telescope optics and detectors. All of these make the interpretation of what is normally a one or two temperature spectral fit to the data difficult to relate to the underlying physical conditions, and conversely, make it difficult to predict the {\\em observable} properties of a model or simulation. To our knowledge there has been no study of the influence of the complexities mentioned above on the best-fit properties of a spectral fit to the observable X-ray data, and in particular not for wind blown bubbles. As we shall show, the combination of the physical effects above and the properties (and limitations) of real X-ray observatories, can significantly affect the results of simple spectral fitting. Previous authors (Weaver \\etal 1977; Zhekov \\& Perinotto 1996) have calculated theoretical X-ray spectra from their 1-D models, but did not consider particular instruments or fit models to those spectra. In general only X-ray luminosities are calculated (\\eg Volk \\& Kwok 1985; Mellema \\& Frank 1995; Garcia-Segura \\& Mac Low 1995). The aim of this paper is to introduce a method of analysing numerical simulations in the same way as actual X-ray observations are analysed, \\ie predict the {\\em observable} X-ray properties. This method can be applied to a wide range of phenomenon where X-rays are important, from wind-blown bubbles around WR stars and PN, through the larger bubbles around clusters of massive stars to starburst-driven galactic winds. We simulate a wind blown bubble using a 2-dimensional hydrodynamic code, concentrating on the properties of the hot X-ray emitting gas. The hydrodynamic model is used to generate artificial X-ray spectra and images, in particular simulated {\\it ROSAT} spectra. We then analyse these spectra in the same way as real {\\it ROSAT} spectra would be, in order to determine what the observationally determined properties of the bubble would be, and how those relate to its actual structure. This synthesis is necessary to a) determine the physical processes that are observationally important, and b) {\\em allow a direct comparison between observation and theory}. Our model of a wind-blown bubble is deliberately chosen to be the simplest applicable model with an analytic solution, in order to simplify the analysis of our results, and avoiding added complications that a more physically accurate model of a wind blown bubble (\\eg Garc\\'{\\i}a-Segura, Mac Low \\& Langer 1996) would introduce into the interpretation of our results. Later papers will consider more realistic models, with additional physics such as time varying energy and mass injection rates, along with spatial variation of the X-ray properties. This will be necessary to understand the properties of the extended emission from galactic winds (see for example Strickland, Stevens and Ponman 1997). In Section~\\ref{sec:num_method} we describe the numerical code used to produce the results shown in Section~\\ref{sec:results}. Section~\\ref{sec:disc} discusses the implications of these results, and we briefly sum up in Section~\\ref{sec:conclusions}. \\begin{figure*} \\vspace{16.0cm} \\special{psfile=fig1_top.eps hoffset=0 voffset=-30 hscale=90 vscale=90 angle=0} \\special{psfile=fig1_bot.eps hoffset=0 voffset=-240 hscale=90 vscale=90 angle=0} \\caption{Logarithm of the gas number density during the simulation at $t = 3500$, $7700$, $10170$ and $14630 \\yr$. At $t = 3500 \\yr$ the bubble has suffered no significant radiative energy loss. Shell collapse is underway at $7700 \\yr$, approximately the time of maximum soft X-ray luminosity, and has just finished at $10170 \\yr$. The bubble then enters the self-similar phase, its properties at $t = 14630 \\yr$ being typical of this stage.} \\label{fig:dens_4t} \\end{figure*} ", "conclusions": "\\label{sec:conclusions} We have shown that in order to compare X-ray observations to theory, it is necessary to consider the {\\em observable} X-ray properties of the theory. The results of a spectral fit are a complex function of the the density and temperature distributions of the source, absorption, the properties of the detector used and the spectral fitting procedure. As such they should not be considered as ``real'' values, but as characteristic values, and specific to the instrument used. The normal method of fitting a simplistic model to the observed data, and then treating the best-fit parameters as the real properties can easily give answers an order of magnitude out from the truth. This technique will allow the first direct comparison between observation and theoretical models of superbubbles and starburst driven outflows.\\\\ We would like to thank Trevor Ponman, Robin Williams and the referee for constructive criticism. DKS and IRS acknowledge financial support from PPARC. This work was performed on the Birmingham node of the {\\sc Starlink} network." }, "9803/astro-ph9803229_arXiv.txt": { "abstract": "Recently gathered observational data on a sample of Type Ia Supernovae (SNe~Ia) reveal a wide distribution of expansion velocities of the Fe cores, measured from the width of the nebular lines. Moreover, the velocity appears to correlate with the luminosity decline rate after maximum light, $\\Delta m_{15}(B)$. Since it has been shown that for SNe~Ia $\\Delta m_{15}(B)$ correlates with the absolute magnitude at maximum, this then implies a relation between the expansion velocity of the Fe nebula and the luminosity at maximum. Physically, the maximum luminosity is related to the mass of synthesized $^{56}$Ni, whereas the $FWHM$ of the lines is related to the kinetic energy of the ejecta. Our finding constitutes observational proof of the theoretical prediction that the two quantities have to be related. ", "introduction": " ", "conclusions": "" }, "9803/physics9803022_arXiv.txt": { "abstract": "A fully nonlinear, time-asymptotic theory of resonant particle trapping in large-amplitude quasi-parallel Alfv\\'en waves is presented. The effect of trapped particles on the nonlinear dynamics of quasi-stationary Alfv\\'enic discontinuities and coherent Alfv\\'en waves is highly non-trivial and forces to a significant departure of the theory from the conventional DNLS and KNLS equation models. The virial theorem is used to determine the time-asymptotic distribution function. ", "introduction": " ", "conclusions": "" }, "9803/astro-ph9803159_arXiv.txt": { "abstract": "We present a high-resolution spectrum of the high redshift, lensed quasar Q1208+1101, obtained with the echellette spectrograph on the Multiple Mirror Telescope. We examine the new and published spectra and provide an updated list of high-confidence metal-line absorption systems at $z=1.1349, 2.8626, 2.9118, 2.9136, 2.9149$. Combining this with a simple model of the gravitational lens system allows us to constrain the possible lens redshifts. The high-redshift ($z > 2.5$) and low-redshift ($z < 0.4$) candidates can be ruled out with high confidence. The current spectra effectively probe about 40\\% of the redshift range in which the lens is expected. In that range, there is only one known metal-line absorption system, an MgII absorber at $z=1.1349$. We consider the possibility that this system is the lensing galaxy and discuss the implied parameters of the galaxy. ", "introduction": "The bright, high redshift (z=3.815) radio quiet quasar Q1208+1011 has been identified as a gravitational lens by Maoz et al. (1992) and Magain et al. (1992). The lens consists of two components (V=18.3 and 19.8 mag, Bahcall et al. 1992) separated by 0.\\arcsec 47, with a 4:1 intensity ratio. The FOS HST spectra (Maoz et al., 1992) show that both components have the same redshift and similar spectra. There are three key aspects in studying gravitational lenses: 1) understanding the lens geometry; 2) understanding the properties of the lensing galaxy; and 3) understanding the properties of the background source. Determining the amount of magnification allows us to understand the intrinsic quasar emission. Given a limiting observed magnitude, lensing allows us to probe both to lower intrinsic luminosities (at a certain redshift) or to higher redshifts (at a certain luminosity). Q1208+1011 is apparently an extremely high luminosity source with an observed optical luminosity of $\\sim 10^{48}$~ergs~s$^{-1}$. The true intrinsic luminosity is likely to be much lower, which affects the modeling and influences the parameters of the quasar models such as required black hole mass or accretion rates (Czerny 1994, Antonucci 1994, Siemiginowska et al 1996). Precise lens modeling and evaluation of the quasar magnification requires detailed information about the lens, including its exact position relative to the quasar images, its morphology or mass distribution, and its redshift (Kochanek 1991). The SIS lens model predicts an average magnification of about 4 (Turner et al. 1984), however we cannot give a correct value for Q1208+1011 until the lens is detected. Bechtold (1994) investigated the proximity effect in the spectra of Q1208+1101 and concluded that the data were consistent with a magnification of 1. Fontana at al. (1997) give a factor of 20 magnification for Q1208+1101 based on high resolution Lyman alpha forest data. Lens detection combined with the proximity effect could give stronger constraints on the magnification factor, and therefore allow more accurate modeling of this very luminous source. Thus far the lensing galaxy for Q1208+1011 has not been directly imaged, consistent with the expectation that it is 4-6 magnitudes fainter than the quasar (Bahcall et al. 1992, Kochanek 1991). The small separation indicates that a galaxy at relatively high redshift, $z \\ga 0.5$, is likely responsible for the lensing (Turner et al. 1984). For this system and others with suspected high-redshift lensing galaxies, it may be possible and even necessary to identify the lens by its {\\em absorption} properties, rather than by its emission. With few exceptions, galaxies within $\\sim 30 h^{-1}$\\,kpc of the quasar line of site cause MgII or CIV absorption (Steidel 1997; Steidel 1993), so one would expect an metal-line absorption system at the lens galaxy redshift. For Q1208+1011, the only published analysis of possible lens redshifts based on absorption lines has been by Magain et al. (1992). They re-analyzed the absorption line data presented by Steidel (1990) and suggested at least 18 possible metal-line absorption systems, spanning redshifts from 0.3741 to 2.9157, with the majority in the range $2.5 < z < 3.1$. They proposed that the most likely lens system was at redshift 2.9157 and derived a corresponding mass estimate for the lens. However, lens models indicate that such a high redshift location is highly unlikely (see Section~\\ref{sec:lens_model}). Furthermore, most of low redshift identifications were based on just two doublet lines within the Ly-$\\alpha$ forest, whereas Bechtold and Yee (1995) have shown that the false detection rate for doublets in the forest is quite high (see also Section~\\ref{sec:abs_lines}). To constrain the lens redshift more reliably in Q1208+1011, we obtained a high-resolution spectrum in March 1996 with the echellette spectrograph on the Multiple Mirror Telescope (MMT). In this paper we examine the new and published spectra and provide an updated list of high-confidence metal-line absorption systems (Section~\\ref{sec:abs_lines} and Section~\\ref{sec:lens_model}). We then combine this with a simple model of the gravitational lens system to constrain the possible lens redshifts (Section~\\ref{sec:lens_model}). We show that the high-redshift ($z > 2.5$) and low-redshift ($z < 0.4$) candidates proposed by Magain et al. can be ruled out with high confidence, and that the current spectra effectively probe about 40\\% of the redshift range in which the lens is expected. In that range, there is only one known metal-line absorption system, an MgII absorber at $z=1.1349$. In Section~\\ref{sec:discussion} we consider the possibility that this system is the lensing galaxy and discuss the implied parameters of the galaxy. We also calculate the expected galaxy IR luminosity. ", "conclusions": "Because the separation between the two quasar images is only 0.\\arcsec 47, the mass of a normal galaxy is adequate to produce the lensed images. Assuming the Singular Isothermal Sphere (SIS) model for the lensing galaxy we can estimate the velocity dispersion and the enclosed mass for a given redshift. In the redshift range which we have searched, there is only one candidate lens redshift, the $z=1.1349$ MgII system. However, since there is a significant probability that this is not the lens redshift, we also calculate the parameters for several other interesting redshifts: $z=0.4$ low-redshift case (low end of the 90\\% probability range); $z=2.4$ high-redshift case (high end of the 90\\% probability range); and $z=2.9$ C\\,IV case, corresponding to the known C\\,IV absorption systems. In all the calculations below, we assume $\\rm \\Omega_0=0.1$ and $\\rm H_0=100h~km~sec^{-1} Mpc^{-1}$. The mass of the galaxy can be obtained from: $$ M \\sim {4 \\theta^2 \\over 9} \\, {D_l D_s \\over D_{ls}}$$ \\noindent where $\\theta$ is the image separation, M is mass of the lens, $D_l, D_s, D_{ls}$ are the angular diameter distances to the lens, to the quasar and between the lens and the quasar respectively (see the review by Blandford \\& Narayan, 1992). The corresponding velocity dispersion is related to the image separation by: $$ \\theta = 4 \\pi {\\sigma^2 \\over c ^2} \\, {D_{ls} \\over D_s} = 2.6 \\arcsec \\sigma ^2 _{300} {D_{ls} \\over D_s} $$ \\noindent where $\\sigma = 300 \\times \\sigma _{300} $~km~s$^{-1} $ is a velocity dispersion. Table 2 contains the calculated mass and velocity dispersions for the four considered lens redshifts. The main uncertainty on the mass is related to cosmology and the uncertainty on the Hubble constant. The required mass ($\\sim 2.8 \\times 10^{11} M_{\\odot}$) and the velocity dispersion ( $\\sim 202$~km~s$^{-1}$ for the lens at $z=1.1349$ are quite typical of normal galaxies. We believe the Mg II absorption system at z=1.1349 is a strong candidate to be the lensing galaxy. Absorption of the kind and strength we see is, with few exceptions, associated with a galaxy within $\\sim30 h^{-1}$\\,kpc of the line of sight (Le Brun, et al. 1995; Steidel, 1993). This implies that there is a galaxy within $\\sim 4$\\arcsec\\ of the Q1208+1011 pair. Additionally, only in rare cases is there a galaxy within $\\sim 30h^{-1}$\\,kpc which {\\em does not} cause Mg\\,II absorption (Steidel 1993). Thus far the lensing galaxy has not been detected -- pre-Costar HST imaging with the PC (Bahcall et al. 1992) limits the galaxy to have V$> 20.7$ if it is more than 0.\\arcsec 5 from the brighter component, or V$>19$ if the galaxy is between the two images. Given the mass estimate for this system, we can predict its brightness using the Tully-Fisher and Faber-Jackson relations. If the galaxy is a disk system, the velocity dispersion implies that it is about 1 magnitude brighter than $L^*$; if it is an elliptical, it is 0.3 magnitudes fainter than $L^*$. Figure 5 shows predicted lens galaxy magnitudes in HST $V$ (F555W) and $H$ (F160W) bands. The luminosities were estimated by combining an SIS lens model with the Faber-Jackson or Tully-Fisher relation, and the magnitudes were then estimated by applying $K$ and evolutionary corrections computed from the spectral evolution models of Bruzual \\& Charlot (1993). (See Keeton, Kochanek \\& Falco 1997 for details.) The predicted apparent magnitude in the visible is V=24.1-25.4 (see Figure~\\ref{fig:prob_dist}), much fainter than the limit $V\\sim20.7$ placed by Bahcall et al (1992) from the pre-Costar PC on HST. The predicted near-IR magnitude is $H\\approx19.2-20.6$ (see Figure~\\ref{fig:prob_dist}), or $K\\approx20.2-21.6$. This is near the faint end of the range of luminosities of galaxies selected by the presence of Mg\\,II absorption and described by Steidel \\& Dickinson (1995). They presented data which showed that, for 5 Mg\\,II systems with $1.0 < z < 1.2$, the galaxy causing the absorption had K magnitude between 18.5 and 20.0. We have simulated NICMOS observations to determine whether such a galaxy will be easily visible. We assume the galaxy is centered between the quasar images, synthesize a test image, and remove the quasar images using a synthesized point-spread-function. We find that a four-orbit exposure with the low background H-band (F160W) filter might give a detection with sufficient signal to estimate a lens model and the corresponding magnification. A single-orbit exposure such as the one planned for Cycle~7 (Falco et al.) \\footnote{Preliminary NICMOS images of Q1208+1011 have recently become available on the CASTLE Web page: {\\texttt http://cfa-www.harvard.edu/glensdata/1208.html}. The galaxy is not apparent in the image consistent with the predicted magnitude.} migth detect the core of the galaxy but will not likely trace the profile very far beyond the quasar image. We note that the image separation of 0.$\\arcsec$47 corresponds to $\\sim 3h^{-1}$~kpc, slightly smaller than typical scale lengths and effective radii of $L^*$ galaxies. \\bigskip We have combined our high-resolution spectra of the metal-line absorption systems towards the lensed quasar Q1208+1101 with gravitational lensing models. We find the MgII absorber at z=1.1349 to be a plausible candidate for the lensing galaxy." }, "9803/astro-ph9803190_arXiv.txt": { "abstract": "We report the detection of luminous extended X-ray emission in NGC 6240 on the basis of \\ros HRI observations of this ultraluminous IR galaxy. The spatial structure and temporal behavior of the X-ray source were analyzed. We find that $\\ga 70\\%$ of the soft X-ray emission is extended beyond a radius of 5\\arcsec. Strong emission can be traced out to a radius of 20\\arcsec~ and weaker emission extends out to $\\sim$50\\arcsec. With a luminosity of at least $L_{\\rm x} \\simeq$ 10$^{42}$ erg/s this makes NGC 6240 one of the most luminous X-ray emitters in {\\em extended} emission known. Evidence for a nuclear compact variable component is indicated by a drop of 32\\% in the HRI count rate as compared to the PSPC data taken one year earlier. No short-timescale variability is detected. The HRI data, which represent the first high-resolution study of the X-ray emission from NGC 6240, complement previous spectral fits to \\ros PSPC data that suggested a two-component model consisting of thermal emission from shocked gas immersed in a starburst wind plus a powerlaw source attributed to scattered light from an obscured AGN. We discuss several models to account for the extended and compact emission. Although pushed to its limits the starburst outflow model is tenable for the essential part of the {\\em extended} emission. For the AGN-type component we propose a model consisting of a near-nuclear `warm scatterer' that explains the apparent fading of the X-ray flux within a year as well as the strong FeK$\\alpha$ complex seen in an \\asca spectrum. ", "introduction": "With a far-infrared luminosity of $\\sim 10^{12} L_\\odot$ (Wright et al.\\ 1984) and a redshift of $z=0.024$, NGC 6240 is one of the nearest members of the class of ultraluminous infrared galaxies (hereafter ULIRG). The basic, as yet unsolved, enigma of these objects is the nature of the primary power source that has to be situated inside the central few arcseconds (Wynn-Williams \\& Becklin 1993). An amount of $\\sim10^{10} M_{\\sun}$ of cold molecular gas (e.g. Solomon et al. 1997), a record 2.121$\\mu$m-line luminosity from shocked `warm' H$_2$ (e.g. van der Werf et al. 1993), earthbound IR spectra (e.g. Joseph \\& Wright 1985, Rieke et al. 1985, Schmitt et al. 1996), MAMA (Smith et al. 1992) and HST (Barbieri et al. 1993) observations, and recent ISO-SWS spectra (Lutz et al. 1996) all point towards the presence of hidden prodigious star formation after onset of a galactic collision, which could be responsible for the FIR power. Heckman et al. (1987, 1990) found indications for superwind and supershell activity, a well-known result of strong starbursts. However, the smallness of the recombination line flux (de Poy et al. 1986), the detection of a high-excitation component in HST images (Barbieri et al. 1995, Rafanelli et al. 1997) and general arguments valid for ULIRGs as a class (e.g. Sanders et al. 1988) suggest that a dust-shrouded AGN contributes significantly to the heating of the dust that emits the FIR radiation. The unambiguous detection and investigation of an AGN in NGC 6240 and other interacting ULIRGs would be of prime importance for our understanding of the formation and evolution of AGN in general. It has been proposed that starbursts are the germ cell for the formation of AGN (Weedman 1983, Barnes \\& Hernquist 1991, Mihos \\& Hernquist 1996) and interaction may provide the triggers and fuel for both kinds of activity (see Sanders \\& Mirabel 1996 for a recent review). A large fraction of ULIRGs indeed turned out to be interacting systems (e.g. Andreasian \\& Alloin 1994, Clements et al. 1996). As outlined above, there are indications for a starburst in NGC 6240, but what would be the best evidence for a hidden AGN? The far-infrared emission is reprocessed black-body like radiation containing no direct clue on the nature of the primary source. Near-IR and mid-IR line spectra provided signatures for a red giant population and a younger burst. A few high-excitation features in IR spectra and in optical HST narrow-band images could be due to an AGN but not necessarily. The optical emission-line spectrum (Fosbury \\& Wall 1979, Zasov \\& Karachentsev 1979, Fried \\& Schulz 1983, Morris \\& Ward 1988, Keel 1990, Heckman et al. 1987, Veilleux et al. 1995, Schmitt et al. 1996) is dominated by LINER-like line ratios over the central $\\sim10$ kpc. Its large extent and little variation in excitation tracers is more easily attributed to shock-heating rather than to a central photoionizing AGN continuum. X-rays are an important tool for studying both, an AGN as well as starburst components. In the \\ros band, AGN tend to be dominated by strong powerlaw (hereafter PL) emission while starbursts can usually be represented by thermal spectra. In a recent analysis of \\ros PSPC spectra from NGC 6240, Schulz et al. (1998; hereafter paper I) found good fits by either a single thermal Raymond-Smith (hereafter RS) spectrum with $L_{\\rm 0.1-2.4 keV} = 3.8\\,10^{43}$ erg/s (for a distance of 144 Mpc) or a hybrid model consisting of 80\\% PL plus 20\\% thermal RS (dubbed 0.8PL+0.2RS below) contributions and a total luminosity of $5.2\\,10^{42}$ erg/s. Since the spectral shape with PSPC resolution is not sufficiently distinctive the luminosity information was taken as an additional constraint. Due to the unprecedented high luminosity of the single RS model and additional severe difficulties to explain it in terms of scalable superbubble models, the hybrid model was favored. This is also supported by the {\\asca} detection of a strong FeK$\\alpha$ line in the X-ray spectrum of NGC 6240 (Mitsuda 1995; the same data indicate further emission lines around 1--2 keV). The powerlaw was attributed to the electron scattered X-ray flux from a hidden AGN so that an AGN-plus-starburst scenario was proposed for the ultimate power source of NGC 6240 (paper I). The deep HRI observations which are discussed below represent the first high spatial resolution study of the X-ray emission from NGC 6240. They allow to trace the emission from a thermal starburst source that is expected to be appreciably spatially extended while an AGN-induced powerlaw source should be much more compact unless there is extensive large-scale scattering. Further, they provide information on the long- and short-term X-ray variability of the source. Luminosities given below are calculated assuming a distance $d = 144$ Mpc of NGC 6240. This yields a scale perpendicular to the line of sight in which 1\\arcsec~ corresponds to 700 pc in the galaxy. ", "conclusions": "We detected luminous extended X-ray emission in NGC 6240 in \\ros HRI data. At the given spatial resolution the source looks nearly spherical and contains its most significant emission within a radius of 20\\arcsec~(or 14 kpc for a distance d=144 Mpc) where the total 0.1--2.4 keV X-ray luminosity amounts to at least $\\sim 10^{42}$ erg s$^{-1}$. At the epochs of the observations at most 40\\% of this luminosity arises within the innermost region of 5\\arcsec~radius. The extended emission can be consistently described by crude supershell models thereby explaining it as the result of a super-starburst with a total luminosity close to $10^{12} L_{\\sun}$. The presence of an additional compact AGN component is in X-rays indicated by (i) a decrease in the count rate between Feb. 1993 (last PSPC observation) and Feb. 1994 (first HRI observation) by 32\\%, (ii) a probable powerlaw component necessary to fit PSPC spectra and (iii) a strong FeK$\\alpha$ complex detected in \\asca spectra. We propose a model in which near-nuclear warm gas ionized by the AGN powerlaw continuum emits FeK$\\alpha$ which is seen superposed on the reflected continuum. Both components, the starburst as well as the AGN provide enough power to explain the luminous FIR emission in NGC 6240 and it seems that both contribute with comparable strength." }, "9803/astro-ph9803049_arXiv.txt": { "abstract": "s{Several conclusions have been reached over the last few years concerning high-redshift galaxies: (1)~The excess of faint blue galaxies is due to dwarf galaxies. (2)~Star formation peaks at redshifts $z\\approx 1-2$. (3)~It appears to occur piecemeal in any given galaxy and there is no evidence for starbursting throughout a large $\\sim 10\\kpc$ galaxy. (4)~There is significant and sharp diminution in the number of $L_\\star$ spiral galaxies at redshifts $1 10^{51}\\,$erg~s$^{-1}$. An impulsive event releasing this amount of energy in a compact region naturally leads to a fireball and thence to a relativistic blast-wave. Blast-wave models for gamma-ray bursts have been examined previously in the literature in several contexts. The most extensive body of work on this topic has been produced by \\Meszaros, Rees and collaborators (e.g., \\Meszaros\\ \\& Rees 1992a; Rees \\& \\Meszaros\\ 1992; \\Meszaros, Laguna \\& Rees 1993; Wijers, Rees \\& \\Meszaros\\ 1997; Panaitescu \\& \\Meszaros\\ 1998a, 1998b; et al.). Other recent papers include Sari \\& Piran (1995), Vietri (1997), Waxman (1997), and Katz \\& Piran (1997). The basic fireball/blast-wave model consists of some triggering event---either coalescing neutron stars or black holes (\\Meszaros\\ \\& Rees 1992b) or the collapse of a massive star (Paczy\\'nski 1998) or a failed type II supernova (Woosley 1993)---depositing a large amount of energy, $E_0 \\sim 10^{51}$--$10^{55}$~ergs, in a small region with radius $r_0 \\sim 10^6$--$10^7$~cm. Because of the unavoidable presence of baryonic material, it is expected that the initial fireball energy will be transformed into kinetic energy of these baryons rather than escape as radiation. This material expands until the internal motions of the baryons become sub-relativistic in the co-moving frame of the material, at which point it forms a cold shell with bulk Lorentz factor $\\G_0 \\simeq E_0/M_0 c^2$, where $M_0$ is the rest mass of the contaminating baryons. This shell continues to expand freely into the surrounding ambient medium until the integrated momentum impulse upon the shell by the swept-up matter is equal to the rest mass of the original material, $\\G_0 4\\pi r_d^3 \\rho_\\ext /3 \\approx M_0$ where $\\rho_\\ext$ is the mass density of the external medium. This defines the so-called deceleration radius $r_d$ (Rees \\& \\Meszaros\\ 1992). Beyond this radius, the shell can no longer be regarded as freely expanding, and the bulk kinetic energy of the blast-wave begins to be reconverted into internal energy. If this internal energy is radiated promptly, then the deceleration of the blast-wave shell can be approximately described by $\\G(r) \\propto r^{-3}$, and the expansion is said to be in the radiative regime. On the other hand, if the internal energy is radiated on a time scale which is long compared to the expansion time scale, then the expansion is in the non-radiative regime, and $\\G(r) \\propto r^{-3/2}$. In either case, for large initial Lorentz factors, $\\G_0 \\sim 10^2$--$10^3$, relativistic effects (e.g., Rees 1966) compress the time scale for the radiation such that the bulk of the blast-wave energy is emitted in the first tens of seconds in the observer's frame following the initial detonation event, thus producing the observed gamma-ray burst. Several authors have pointed out that well after the prompt gamma-ray burst event the blast-wave shell will continue to decelerate and radiate (Vietri 1997; Waxman 1997). The recent detections by the X-ray, optical and radio communities of the aforementioned fading X-ray, optical and radio counterparts within the error boxes of GRBs appear to support this picture. Furthermore, model estimates of the temporal decay of these transients yield time-dependencies which agree with those observed, $F_\\nu \\sim t^{-1}$ (Wijers et al.\\ 1997); and for the one of the bursts for which optical data are available (GRB~970508), the optical spectral index is consistent with synchrotron emission from a power-law distribution of electrons, $dN/d\\g \\propto \\g^{-p}$ with $p \\simeq 2$--2.3 (Djorgovski et al.\\ 1997) indicating, for example, a shock-accelerated electron population. Despite the successes of the blast-wave model in accounting for the prompt burst properties and its prediction of fading afterglows, several important theoretical questions must be addressed in order to have a reasonably complete model: \\begin{itemize} \\item What is the nature of the coupling between the electrons, protons and magnetic field (Panai\\-tescu \\& \\Meszaros\\ 1998b)? To what extent can these components be in equipartition given that only the electrons can efficiently radiate away their energy? \\item How does the magnetic field change as the blast-wave decelerates? If it is initially formed through equipartition processes, but is not strongly coupled to the non-thermal electron energy density, how does it evolve? \\item What is the nature of the acceleration mechanism? Are the electrons energized by repeated diffusion across the shocks themselves or by gyroresonant scattering with disturbances in the post-shock turbulent MHD fluid? \\item What is the proper form for the injected electron energy distributions? Is it well described by a typical power-law spectrum? If so, what determines the characteristic energies of the particles? \\end{itemize} In order to begin to address these questions, it is worthwhile to go beyond the simple, though useful, back-of-the-envelope estimates which have generally prevailed in the literature thus far. In this paper, we present a detailed calculation of the blast-wave deceleration and the evolution of the magnetic fields and electron distributions under various assumptions relevant to the above issues. We compute model light curves and spectra and use the available afterglow data to help discriminate between the various options. The format of the paper is as follows: In \\S~2, we describe the basic dynamics of an impulsively driven blast-wave which decelerates by accretion of ambient material. In \\S~3, we discuss the physical processes responsible for producing the observed radiation including prescriptions for magnetic field generation, the formation of non-thermal particle distributions and the relevant radiation processes. The numerical procedures for computing the deceleration of the blast-wave and the integration of the blast-wave shell emission are described in \\S~4. Model spectra and light curves for gamma-ray bursts are presented and analyzed in \\S~5. Lastly, in \\S~6, we discuss these results, explore prospects for further research and present our conclusions. ", "conclusions": "In this paper, we have attempted a more realistic calculation of the dynamics and synchrotron and synchrotron self-Compton emission for the blast-wave model of gamma-ray bursts. By matching the detailed characteristics of burst spectra, we have found relations (eqs.~\\ref{Gamma_limit} \\&~\\ref{xi_B_limit}) which place constraints on magnetic field strengths and bulk Lorentz factors. If these relations are to be believed, then burst data can have a significant impact on models of magnetic field generation in turbulent plasmas. The apparent deficiencies of this calculation point towards areas of further research. In particular, the detailed temporal structure of individual bursts is not explicitly dealt with in this model. In the context of external shocks, it may be due to inhomogeneities in the external medium, or fluctuations in the electron injection and/or magnetic field equipartition parameters (Panaitescu \\& \\Meszaros\\ 1998a). For $s > 3$, the generalized expression for the luminosity for energies $\\e \\ge \\e_\\peak$ (eq.~\\ref{peak_luminosity}) is \\begin{equation} \\e^2 \\frac{dN}{d\\e dt} = 2\\pi m_p c^2 \\xi_e r^2 n_\\ext(r) \\G^2(r) (\\e/\\e_\\peak)^\\lambda \\end{equation} where $\\lambda = (2-s)/2$ applies for relatively strong magnetic fields $\\xi_B \\ga 10^{-4}$ when the electrons just above the break are efficiently cooled and $\\lambda = (3-s)/2$ applies for relatively weak fields and uncooled electrons. From this expression we see that any burst light curve substructure must be due to variations in $\\xi_e$ and $n_\\ext$, and indirectly, due to variations in $\\xi_B$ through $\\e_\\peak$ (eq.~\\ref{peak_energy}). Our treatment of the dynamics also ignores the structure of the shock region itself, and our approach essentially only considers the emission from the forward shock and neglects the reverse shock. Panaitescu \\& \\Meszaros\\ (1998a) have performed calculations similar to our own, but from a hydrodynamical perspective, and found that the reverse shock only makes a significant contribution to the emission at optical and UV energies. Therefore, neglecting the reverse shock should not affect our results for the gamma-ray emission, but it could have a significant impact on the optical and radio afterglow emission. We also neglect the thickness, $\\Delta r$, of the shock shell in integrating the emission for a given observer time $\\dt$. This should not be important at early times when $\\Delta r = r_0/\\G_0^2$ (in the lab frame), but it could affect the afterglow emission at late times. It is unlikely that the blast-wave itself is spherical. If the initial fireball is created by the coalescence of two compact objects, then the orbital plane defines a natural axis of symmetry along which the blast-wave will propagate (\\Meszaros\\ \\& Rees 1992b). This sort of asymmetry could be accounted for in our model by a non-unity collimation factor, $f_b$. Furthermore, if the observer line-of-sight does not lie within the opening angle of the blast-wave cone, then other effects due to relativistic beaming and the gradual deceleration of the shock front would be introduced. In this respect, highly anisotropic blast-waves would share properties with relativistic jets in blazars. This analogy can be take even further by noting the similarity of the burst spectra we derive compared to that of gamma-ray blazars. Like our model spectra, the spectral energy distributions (SEDs) of these objects tend to have two peaks, one in the UV--X-ray range and one at gamma-ray energies. If the lower peak in blazar SEDs is due to synchrotron emission and corresponds to the $\\sim 1$~MeV peak in gamma-ray bursts, we can apply a similar analysis as we have discussed above to derive bulk Lorentz factors and equipartition parameters for blazars. In particular, the recent ASCA observations of Mrk~421 (Takahashi et al.\\ 1996) provide sufficient information to get actual values rather than simply upper or lower limits. Using the light curves of Mrk~421 measured in different X-ray energy bands, Takahashi et al.\\ (1996) performed a cross-correlation analysis and found that the longer relative time lags of the lower energy data versus the higher energy data are consistent with synchrotron cooling of the underlying electron distribution. Several authors have noted this effect and have calculated this temporal dependence for the cases of bursts and blazars (e.g., Tashiro et al.\\ 1995; Tavani 1996; Dermer 1998). Takahashi et al.\\ use the TeV variability time scale (Kerrick et al.\\ 1995) to estimate a Doppler factor and find $\\D = 5$ (cf.\\ Takahara 1994). Using this estimate and their time lag measurements, they derive a magnetic field of $B = 0.2$~G. From non-simultaneous data (Shrader \\& Wehrle 1997), the synchrotron portion of the SED of Mrk~421 peaks at about $\\sim 130~$eV. Using \\begin{equation} \\e_\\peak = \\frac{B}{B_{\\rm crit}} \\g^2 {\\cal D} \\end{equation} (cf.\\ eq.~\\ref{peak_energy}), and $\\g = (m_p/m_e)\\G$, we find $\\G \\approx 60$ and an observer angle $\\theta \\approx 5^\\circ$. We also obtain an equipartition field strength of $B_{eq} \\approx 10 n_1^{1/2}$~G implying an equipartition parameter of $\\xi_B \\sim 10^{-2}$. Although the above value for the bulk Lorentz factor is substantially larger than the mean value of $\\langle \\G \\rangle \\sim 10$ found by applying the beaming model to a sample of radio-loud objects (Urry \\& Padovani 1995), its large value may indicate the special nature of gamma-ray loud blazars which are characterized not only by small observing angles but also by larger than typical bulk Lorentz factors. Despite the crudeness of this calculation, it illustrates the potential applicability of this sort of analysis to blazars as well as bursts." }, "9803/astro-ph9803175_arXiv.txt": { "abstract": "We report the detection of four images in the recently discovered lensed QSO RX~J0911.4+0551. With a maximum angular separation of $3.1$\\arcsec, it is the quadruply imaged QSO with the widest known angular separation. Raw and deconvolved data reveal an elongated lens galaxy. The observed reddening in at least two of the four QSO images suggests differential extinction by this lensing galaxy. We show that both an ellipticity of the galaxy ($\\epsilon_{\\rm min}=0.075$) and an external shear ($\\gamma_{\\rm min}=0.15$) from a nearby mass has to be included in the lensing potential in order to reproduce the complex geometry observed in RX~J0911.4+0551. A possible galaxy cluster is detected about 38\\arcsec\\, from RX~J0911.4+0551 and could contribute to the X-ray emission observed by ROSAT in this field. The color of these galaxies indicates a plausible redshift in the range of 0.6-0.8. ", "introduction": "RX~J0911.4+0551, an AGN candidate selected from the ROSAT All-Sky Survey (RASS) (Bade et al. 1995, Hagen et al. 1995), has recently been classified by Bade et al. (1997; hereafter B97) as a new multiply imaged QSO. B97 show that it consists of at least three objects: two barely resolved components and a third fainter one located 3.1\\arcsec\\ away from the other two. They also show that the spectrum of this third fainter component is similar to the combined spectrum of the two bright components. The lensed source is a radio quiet QSO at $z=2.8$. Since RASS detections of distant radio quiet QSOs are rare, B97 pointed out that the observed X-ray flux might originate from a galaxy cluster at $z \\geq 0.5$ within the ROSAT error box. We present here new optical and near-IR high-resolution images of RX~J0911.4+0551 obtained with the 2.56m Nordic Optical Telescope (NOT) and the ESO 3.5m New Technology Telescope (NTT). Careful deconvolution of the data allows us to clearly resolve the object into four QSO components and a lensing galaxy. In addition, a candidate galaxy cluster is detected in the vicinity of the four QSO images. We estimate its redshift from the photometric analysis of its member galaxies. ", "conclusions": "Thanks to our new high-resolution imaging data, the QSO RX~J0911.4+0551 is resolved into four images. In addition, deconvolution with the new MCS algorithm reveals the lensing galaxy, clearly confirming the lensed nature of this system. The image deconvolution provides precise photometry and astrometry for all the components of the system. Reddening in components A2 and A3 relative to A1 is observed from our $U$, $V$, and $I$ frames that were taken within three hours on the same night. The absence of reddening in component B and the difference in reddening between components A2 and A3 suggest extinction by the deflecting galaxy. Note that although our near-IR data were obtained from 15 days to 6 weeks after the optical images, they appear to be consistent with the optical fluxes measured for the QSO images, i.e. flux ratios increase continuously with wavelength, from $U$ to $K$, indicating extinction by the lensing galaxy. We have discovered a good galaxy cluster candidate in the SW vicinity of RX~J0911.4+0551 from our field photometry in the $I$, $J$, and $K$ bands. Comparison of our color-magnitude diagram with that of a blank field (e.g., Moustakas et al. 1997) shows that the galaxies around RX~J0911.4+0551 are redder than field-galaxies at an equivalent apparent magnitude. In addition, the brightest galaxies in Fig.~\\ref{fig:cmd} lie on a red sequence at $I-K\\sim 3.3$, typical for the early type members of a distant galaxy cluster. The two dashed lines indicate our $\\pm0.4$ color error bars at $K\\sim 19$ around $I-K\\sim3.3$. Most of these galaxies are grouped in the region around a double elliptical at a distance of $\\sim38$\\arcsec\\, and a position angle of $\\sim204^{\\circ}$ relative to A1. This can also be seen in Fig.~\\ref{fig:field} which shows a group of red galaxies with similar colors centered on the double elliptical (in the center of the circle). Consequently, there is considerable evidence for at least one galaxy cluster in the field. The redshift of our best candidate cluster (the one circled in Fig.~\\ref{fig:field}) can be estimated from the $I$ and $K$ band photometry. We have compared the $K$-band magnitudes of the brightest cluster galaxies with the empirical $K$ magnitude vs. redshift relation found by Arag{\\'o}n-Salamanca et al. (1998). We find that our cluster candidate, with its brightest $K$ magnitude of about $\\sim17.0$, should have a redshift of $z\\sim0.7$. A similar comparison has been done in the $I$-band without taking into account galaxy morphology. We compare the mean $I$ magnitude of the cluster members with the ones found by Koo et al. (1996) for galaxies with known redshifts in the Hubble Deep Field and obtain a cluster redshift between 0.6 and 0.9. Finally, comparison of the $I-K$ color of the galaxy sequence with data and models from Kodama et al. (1998) confirm the redshift estimate of 0.6-0.8. In order to calculate physical quantities from the model parameters found in section 4, we assume a simple model for the cluster which may be responsible for the external shear. For an isothermal potential, the true shear and convergence are of the same order of magnitude. As the convergence is not explicitly included in the model, the deduced shear is a reduced shear leading to an absolute convergence of $\\kappa = \\gamma/(1+\\gamma) = 0.241$. For a cluster redshift of $z_{\\rm d}=0.7$ and with cosmological parameters $\\Omega=1$, $\\lambda=0$ this corresponds to a velocity dispersion of about $1100\\,\\kms$ if the cluster is positioned at an angular distance of 40\\arcsec\\,. See Gorenstein, Falco \\& Shapiro (1988) for a discussion of the degeneracy preventing a direct determination of $\\kappa$. From the direction of the shear $\\phi$, (see Table ~\\ref{tab:bestmod}) we can predict the position angle of the cluster as seen from the QSO to be $12^\\circ$ or $192^\\circ$. The latter value agrees well with the position of our cluster candidate SW of the QSO images. Note also the good agreement between the position angle $\\thg$ derived from the observed light distribution, and the predicted position angle corresponding to our best fitting model of the lensing potential. Interestingly, this is in good agreement with Keeton, Kochanek \\& Falco (1998) who find that projected mass distributions are generally aligned with the projected light distributions to less than $10^{\\circ}$. The color of the main lensing galaxy is very similar to that of the cluster members, suggesting that it might be a member of the cluster. Using the same model for the cluster as above, assuming the galaxy at the same redshift as the cluster, and neglecting the small ellipticity of $\\epsilon<0.05$, the velocity dispersion of the lensing galaxy can be predicted from the calculated deflection angle $\\alpha_0$ to be of the order of $240\\,\\kms$. Since the galaxy profile is sharp towards the nucleus in $K$, we cannot rule out the possibility of a fifth central image of the source, as predicted for non-singular lens models. Near-IR spectroscopy is needed to get a redshift determination of the lens and to show whether it is blended or not with a fifth image of the (QSO) source. Some 10\\arcsec\\, SW from the lens, we detect a small group of even redder objects. These red galaxies can be seen in Fig.~\\ref{fig:field} a few arcseconds to the left and to the right of the cross. They might be part of a second galaxy-group at a higher redshift, and with a position in better agreement with the X-ray position mentioned by B97. However, since the measured X-ray signal is near the detection limit, and the 1-$\\sigma$ positional uncertainty is at least 20\\arcsec\\,, the X-ray emission is compatible with both the QSO and these galaxy groups in the field. Furthermore, this second group, at $z>0.7$, would most likely be too faint in the X-ray domain to be detected in the RASS. In fact, even our lower redshift cluster candidate would need to have an X-ray luminosity of the order of $\\rm L_{0.1-2.4 \\rm keV}\\sim 7.10^{44} \\rm erg\\,\\rm s^{-1}$ (assuming a 6 keV thermal spectrum, $\\rm H_{0}=50\\, \\rm{km\\,s}^{-1}\\,\\rm Mpc^{-1}$, $\\rm q_{0}=0.5$), in order to be detected with 0.02 $\\rm cts\\,\\rm s^{-1}$ by ROSAT. This is very bright but not unrealistic for high redshift galaxy clusters (e.g., MS~1054-03, Donahue, Gioia, Luppino et al. 1997). RX~J0911.4+0551 is a new quadruply imaged QSO with an unusual image configuration. The lens configuration is complex, composed of one main lensing galaxy plus external shear possibly caused by a galaxy cluster at redshift between 0.6 and 0.8 and another possible group at $z>0.7$. Multi-object spectroscopy is needed in order to confirm our cluster candidate/s and derive its/their redshift and velocity dispersion. In addition, weak lensing analysis of background galaxies might prove useful to map the overall lensing potential involved in this complex system." }, "9803/astro-ph9803205_arXiv.txt": { "abstract": "This paper discusses the properties of scattering--dominated active galactic nuclei (AGN). We define these to be AGN for which the direct line-of-sight to the continuum source is obscured by Compton-thick material. The aim is to construct, for the first time, a model consistent with X-ray line luminosities, line ratios and various luminosity indicators. The \\ASCA\\ spectra of six such sources show several X-ray lines that can be reliably measured, mostly due to highly ionized magnesium, silicon sulphur and iron. These enable us to investigate the physical conditions of the scattering material. The sources show evidence of He-like and H-like iron lines that are likely to be produced in hot (T$\\sim 10^6$ K) photoionized gas. By measuring the EW of the lines, and by constructing a diagnostic line-ratio diagram, we demonstrate that the silicon and magnesium lines are produced by the same gas emitting the highly ionized iron lines. The properties of this gas are rather different from the properties of warm absorbers in type I AGN. Neutral 6.4 keV iron lines are also detected, originating in a different component which can be either Compton-thin or Compton-thick. The best measured iron lines suggest an enhancement of ${\\rm n_{Fe}/n_H}$ by a factor $\\sim 2$ compared to solar, in both the hot and cool Compton-thin components. We further show that in four of the sources, the Fe \\Ka(6.4~keV)/\\Hb\\ ($\\lambda 4861 \\AA$) line ratio is consistent with that predicted for typical narrow line region clouds, and the reddening corrected \\Hb\\ is known, provided the column density is larger than $\\sim 10^{22.5}$ \\cmii\\ , \\aox\\ is smaller than 1.3. For some sources, this is a viable alternative to the commonly assumed Compton thick medium as the origin of the 6.4 keV iron line. {\\it Subject headings:} galaxies:abundances - galaxies:Seyfert - galaxies:active - line:formation - X-ray:galaxies ", "introduction": "The optical and X-ray properties of type II AGN (i.e. those showing prominent narrow emission lines and very faint, if any, broad lines) have been discussed in numerous recent papers that contain the analysis of their morphology, spectrum, geometry and relationship to type I (broad emission line) AGN. Various names, including Seyfert 2, narrow emission line galaxies (NELG), and narrow line X-ray galaxies (NLXGs) have been used to describe these objects. Obviously, there is some subjectivity in the classification of type II AGN leading to the assignment of a more than one ``type'' for some objects which have been studied by several authors. The geometry of the innermost region of such sources is a fundamental, yet still an open issue and the reader is referred to Antonucci (1993), and Mulchaey \\etal\\ (1994) for discussion and references regarding these questions. X-ray observations offer a unique view of type II AGN, since X-rays can penetrate large column densities. This has been a subject of much research, for example, see recent papers by Turner \\etal\\ (1997a,b,1998). These authors found some surprising similarities between the X-ray spectra of type I and type II AGN. In some type II sources, the equivalent width (EW) of the 6.4 keV line is similar to that observed in Seyfert 1s and the line profiles show broad, redshifted wings. In other type II sources, like NGC~1068, EW(Fe \\Ka) is an order of magnitude larger than in type I AGN, suggesting that the line is seen against a reduced continuum, presumably due to obscuration. Evidently, type II AGN fall into at least two X-ray categories; those where the central source is directly observed below 10 keV, and those where it is not. Hereafter we refer to those type II AGN whose 0.5--10 keV spectra are dominated by scattered radiation, ``scattering--dominated AGN'', and they are the subject of this paper. We expect a good, but not necessarily a one-to-one correlation between such objects and those type II AGN who show highly polarized continuum and broad optical/UV emission lines, due to scattering (e.g. Tran 1995). This paper investigates the properties of scattering--dominated AGN through detailed analysis of their 0.5--10 keV spectrum. We address the nature of the scattering medium and try to deduce its level of ionization, column density and covering fraction. We also investigate the metallicity of the gas and compare its properties to the ionized gas in Seyfert 1 galaxies, and to the narrow line region (NLR) gas. The analysis is aimed at a small number of scattering--dominated AGN whose \\ASCA\\ spectra are of a sufficiently high quality to enable the measurement of at least 3 X-ray emission lines. It also suggests several new avenues for future study of such sources in preparation for the coming {\\it AXAF}\\ and {\\it XMM}\\ missions. In \\S2 we discuss the predicted spectra of such sources. In \\S3 we compare predictions to a detailed analysis of the \\ASCA\\ spectra of six such galaxies. In \\S4 we discuss several implications of such a comparison, and implications for the state and the location of the scattering medium and its composition. ", "conclusions": "The following analysis is based on the line fluxes listed in Table 1 as well as on the analysis of the \\ASCA\\ spectrum of NGC~1068 (Netzer and Turner, 1997, Table 1). Obviously, the number of objects, and the number of measurable lines per object, are very small and the information content regarding the group properties is rather limited. The most severe complication in analyzing the \\ASCA\\ spectra is the likely contamination of the nuclear spectrum by extended, non-nuclear emission. This may be the result of hot gas in star forming regions, supernova remnants, or any other gas at T$\\simeq 10^7$~K. The limited \\ASCA\\ spatial resolution (with a half-power diameter $\\sim3$~arcmin), combined with the relative weakness of the scattered X-ray continuum, makes it almost impossible to separate the photoionized gas and hot plasma contributions. Indeed, some sources (e.g. NGC~1068) show clear indication of extended X-ray emission which is most likely due to starburst activity. This issue will not be resolved before {\\it AXAF} observations (with a half-power diameter $<$1~arcsec). Below we comment on the information obtained from the study of the Fe \\Ka\\ complex and address the potential use of diagnostic diagrams in the analysis of the spectrum of scattering--dominated AGN. \\subsection{The \\Ka\\ complex and the iron abundance} Except for NGC~1068, all our measurements of the 6--7 keV complex are somewhat ambiguous since we can not reliably resolve the highly ionized (6.7 and 6.96 keV) Fe \\Ka\\ lines from the neutral \\Kb\\ ($\\sim 7.1$ keV) line. Therefore, the analysis of the high ionization lines pertains to the {\\it combined intensity} of the H-like and He-like iron lines, which was obtained by subtracting the expected \\Kb\\ flux (10\\% the flux of the 6.4 keV \\Ka\\ line) from the total. This makes the combined Fe{\\sc xxv-xxvi} line intensity in NGC~3488 consistent with zero because of the uncertainty on the \\Kb\\ flux (see Table 1) and the ones in Mkn~348 consistent with zero because of the large intrinsic error. The uncertainty in the relative strength of the 6.4 and 6.9 keV component is also affected by the uncertainty in the assumed, scattered broad 6.4 keV line. The galaxies with measurable soft X-ray lines represent two different groups. In one object (NGC~1068), the highly ionized iron lines are comparable in strength to the 6.4 keV line. In three others (Circinus, Mkn~3 and NGC~4388) the combined intensity of the He-like and H-like iron lines is only about 10-15\\% the intensity of the low ionization component. Mkn~348 is possibly an intermediate case but the observational uncertainties are too large to tell. The intensity of the high ionization iron lines in NGC~6240 is unknown, because of their proximity to the strong 7.1 keV absorption feature due to $2 \\times 10^{24}$ \\cmii\\ of neutral absorber. However, a comparison of the 6.4 keV intensity with the soft X-ray lines suggests that this source belongs to the same group as Mkn~3 and Circinus. As for the lines of silicon, magnesium and sulphur, there are only three sources where reliable line ratios can be obtained and they all look quite similar (see below). Regarding the 6.4 keV iron line, a comparison of the measured line intensities with the calculations shown in Figs. 2 and 3, indicate that in all sources where this line is much stronger than the 6.7--7.0 keV complex, the line can not originate in the same component producing the strong magnesium, silicon and sulphur lines. We therefore suggest that in those sources, much of the 6.4 keV line emissivity is due to reprocessing in a large column of low-ionization gas. As for NGC~1068, the 6.4 keV line in this source may be due to warm (T$_e \\simeq 2 \\times 10^5$~K) photoionized gas (Marshall \\etal\\ 1993; Netzer \\& Turner 1997). This idea is in conflict with the Iwasawa \\etal\\ (1997) model (see below). We conclude that in four (perhaps five) of the six sources, the reprocessing efficiency (\\Rf) of the neutral component is much greater (a factor of 5--10) than that of the ionized component. Given the level of ionization of the gas producing the 6.4 keV iron line, and the shape of the ionizing continuum, we can estimate the iron abundance from the observed EW(Fe \\Ka) in two interesting limits. The first corresponds to Compton--thin gas and has been discussed by Krolik and Kallman (1987), Matt \\etal\\ (1996) and others. In this case, assuming negligible resonance line absorption and complete isotropy of the scattered continuum radiation (as appropriate for a case where the scattering medium is viewed at all possible angles), \\begin{equation} {\\rm EW(6.4~keV) \\simeq 3.17 [\\frac {1.11^{1-\\Gamma}}{2+\\Gamma}] [\\frac{F_Y}{0.3}] [\\frac{ n_{Fe}/n_H }{ 4 \\times 10^{-5} }] \\,\\, keV } \\,\\, , \\end{equation} where F$_{\\rm Y}$ is the fluorescence yield (of the order of 0.3 for low ionization iron). In deriving this expression we have adopted the small column density limit which allows us to neglect the absorption of the emitted \\Ka\\ photons on the way out. A much slower (logarithmic) dependence on ${\\rm n_{Fe}/n_H}$ is expected at very large columns. The second case corresponds to Compton--thick gas, such as the walls of the hypothetical nuclear torus. This case has recently been calculated by Matt \\etal\\ (1996, 1997) for a range of metallicities and viewing angles. For a $\\Gamma=1.9$ continuum, larger than solar ${\\rm n_{Fe}/n_H}$ and $\\cos (i)=0.5$, where $i$ is the angle between the line of sight and the axis of the torus, Matt \\etal\\ estimate (see their Fig. 2), \\begin{equation} {\\rm EW(6.4~keV) \\simeq 1.8 (1+0.6 \\log [\\frac{ n_{Fe}/n_H }{ 4 \\times 10^{-5} }] ) \\,\\, keV .} \\end{equation} The range of observed angles can change this value by up to a a factor of 1.5. Thus, for a typical $\\Gamma=1.9$ continuum, the Compton--thin and Compton--thick assumptions result in a factor 2 difference in the estimated iron abundance. The two cases also predict very different 6-10 keV continuum shape since the Compton-thick gas produces a much stronger 7.1 keV absorption edge. Finally, in Compton-thick gas, some 10\\% of the 6.4 keV line intensity is in a broad red wing. Below we discuss the iron abundance under the two different scenarios. While the emphasis is on line intensities, we note that our continuum fits do not require the presence of a 7.1 keV absorption in any source except NGC~6240. The following examples consider two gas components, one represents ``cold'' (small Ux) gas and the other ``hot'' (large Ux) gas. Each scatters the central radiation and produces a scattered continuum. We refer to these as the ``two continuum components''. We first assume both components to be Compton-thin. In estimating the iron abundance from the observed EW, we note that the two continuum components are contributing at 6.4 keV and $\\Gamma$ is not directly measurable since scattering affects the observed continuum shape (\\S2.2). We use the numerical calculations (Fig. 2) and assumed that the iron composition is identical in all components. We also note that resonance absorption is negligible, because iron is in a very low ionization state in the cold component and the column density is large in the hot component. Given these assumptions, we expect each component (hot or cold) to have a similar EW(Fe \\Ka) {\\it relative to its own continuum}. Thus, in those sources where the 6.7--6.96 keV iron lines are much weaker than the neutral 6.4 keV lines, most of the 6.4 keV continuum is due to reflection by the neutral component. Given the assumed continuum shape, this implies an iron over abundance by a factor 2--3 for Circinus and a factor of 1--1.5 for Mkn~3. For NGC~6240 and NGC~4388, we estimate the EW relative to the scattered component by using our best fit model of these source(Fig. 6). According to the model, about half the observed 6.4 keV continuum is due to transmission and the other half due to scattering. We can thus estimate EW(6.4 keV) relative to the scattered component and deduce n$_{\\rm Fe}$/\\nh$\\simeq 1.5-2\\times$solar. Mkn~348 is so different in this respect that we cannot reliably estimate EW(Fe \\Ka) relative to the scattered continuum. The iron abundance in NGC~1068, assuming a Compton-thin gas, has been discussed by Marshall \\etal\\ (1993) and Netzer \\& Turner (1997), and found to be about three times solar. Thus, under the Compton-thin assumption, we have indications of iron overabundance in 5 sources. Regarding the Compton-thick case, the iron abundance inferred from the observed EW(\\Ka\\ 6.4 keV) line is about half the value deduced above, i.e. consistent with solar for all sources. As we show below, the analysis of the highly ionized gas enables an independent check on the iron composition because the medium producing such lines is unlikely to be Compton-thick. \\subsection{Diagnostic diagrams} Diagnostic diagrams, involving various line ratios, have been used to separate Seyfert galaxies from galactic HII regions, and to search for the spectroscopic signature of LINERs (e.g. Baldwin, Phillips \\& Terlavich, 1981). Below we attempt to use the same method in the X-ray domain, in search for the unique signature of scattering--dominated AGN. There are two major differences between our study and the investigation of the optical spectrum of LINERs and HII regions. First, the number of available X-ray lines is very small and we can only measure, reliably, three line ratios and construct two such diagrams. Second, given the gas is photoionized, the X-ray line spectrum is dominated by recombination and not a single, purely collisionally excited line is strong enough to be used. Collisionally excited X-ray lines dominate the spectrum of hot plasmas with ratios vastly different from that expected in photoionized gas. Thus, there is no overlap in properties and line ratio diagrams can either be used for photoionized gas or for hot plasmas. In contrast, lower excitation photoionized nebulae contain a mixture of recombination and collisionally excited lines that provide very useful diagnostics. Thus ratios like \\bOIIIb/\\Hb\\ have been used to derive the level of ionization of the gas and line ratios involving highly excited O$^{+2}$ transitions have been used to investigate the role of shock excited gas in the spectrum of LINERs and Seyfert galaxies (e.g. Ferland and Netzer 1983). This kind of analysis is not yet possible in the X-ray regime. Fig. 7 shows a diagnostic diagram composed of the best observed line ratios in our sample, I(\\SiXIII)/I(\\FeXXV\\ + \\FeXXVI) versus I(\\MgXI)/I(\\SiXIII). The first ratio is a good indicator of regions of large ionization parameters, with electron temperature of the order of 10$^6$ K, and the second measures the conditions in lower ionization gas. Obviously, the excitation and ionization of \\MgXI\\ and \\SiXIII\\ is rather similar and future analysis, based on oxygen and neon lines, will be of greater use. Measurements of the two ratios are available for Circinus and NGC~1068 and an upper limit can be obtained for Mrk~3 (see \\S2). The diagram shows the location of the three objects along with four theoretical curves, this time assuming the L$_{\\rm E}\\propto {\\rm E}^{-0.5}$ continuum. Similar results, with appropriate scaling of Ux, are obtained for the L$_{\\rm E}\\propto {\\rm E}^{-0.9}$ continuum used in most other calculation. The curves are series of increasing Ux for various column density and gas composition. The solid lines are standard composition models for three column densities, \\Ncol=10$^{22.4}$~\\cmii, 10$^{23}$~\\cmii\\ and 10$^{23.3}$~\\cmii. The dotted line is for \\Ncol=10$^{23}$~\\cmii\\ but with ${\\rm n_{Fe}}$/\\nh\\ three times larger. Inspection of the line ratio diagram, and the observed spectra, suggest that: \\begin{enumerate} \\item\tThe observed line ratios cannot be simultaneously obtained in a single temperature collisionally ionized gas. Under such conditions, the temperature required to ionize Fe{\\sc xxv} and Fe{\\sc xxvi} is inconsistent with the observed strength of the silicon and magnesium lines. If all lines are produced in a single component, this gas must be photoionized. Indeed, \\ASCA\\ spectra of starburst galaxies (e.g. Ptak \\etal, 1997) generally show strong silicon and sulphur lines but little or no emission from neutral Fe \\Ka. \\item\tThe inferred ionization parameter, $U_X\\sim 1$, is 3--10 times larger than the ionization parameter of the highly ionized (warm absorber) gas in Seyfert 1 galaxies (George \\etal, 1998). This is in agreement with the finding of Turner \\etal\\ (1997a). We have examined the properties of this gas and found a mean electron temperature of about 10$^6$ K and no noticeable soft absorption features. Such gas, on the line of sight to a typical AGN continuum, with a column density not exceeding 10$^{23}$ \\cmii\\ (the column density of the great majority of warm absorbers in Seyfert 1 galaxies, see George \\etal\\ 1998) would escape detection by \\ASCA\\ type instruments. \\item\tThe diagnostic diagram cannot, by itself, be used to infer the Fe/Si abundance ratio since large column densities mimic the appearance of a small-column with large Fe/Si. As argued in \\S4.1, the analysis of the 6.4 keV lines suggest large ${\\rm n_{Fe}}$/\\nh\\ in all sources if the emitting medium is Compton-thin. If the high ionization components have similar compositions, then according to the diagram, they must have relatively small column densities, perhaps similar to the lowest column shown in Fig. 7. \\item \tThe weakness of \\MgXII\\ and \\SiXIV\\ lines is somewhat surprising. The lines are predicted to be similar in strength to the lower ionization magnesium and silicon lines (Figs. 2 and 3) yet we could only obtain upper limits. The difficulty of detecting the \\SiXIV\\ line may partly be explained by the notorious detector/mirror features in \\ASCA\\ around 2 keV. \\end{enumerate} \\subsection{The scatterer location and the value of \\Rf} So far we have focused on relative line intensities and line-to-continuum flux ratios. These are useful in determining the ionization and composition but do not reveal the reprocessing efficiency, \\Rf, since EW(Fe 6.4 keV) is almost independent of the column density (Fig. 1). \\Rf\\ can not be obtained by comparing the absolute line flux with the intrinsic luminosity since the latter is not known. However, there are several other luminosity indicators at longer wavelength, including the infrared flux and the \\bOIIIb\\ luminosity (e.g. Mulchaey \\etal, 1994, see extensive discussion in Turner \\etal, 1997b,1998), that can be used. Here we chose to use the reddening-corrected narrow \\Hb\\ line as our luminosity indicator. This is similar to the L(\\bOIIIb) method used in Turner \\etal\\ but enables a more direct comparison with the intrinsic ultraviolet luminosity. Measurements of the narrow \\Hb\\ flux for the six sources, as well as for most known type II AGN, are readily available (Mulchaey \\etal, 1994; Polletta \\etal, 1996; and references therein). The above references contain also the measured \\Ha/\\Hb\\ line ratio which we use to correct for reddening and to obtain the intrinsic \\Hb\\ flux. In the following we assume an intrinsic I(\\Ha)/I(\\Hb)=3.0 and a simple galactic type reddening with A$_{\\rm V}$/E$_{\\rm B-V}$=3.1. Reddening corrected \\Hb\\ fluxes obtained this way are listed in Table 1. A word of caution is in order. Applying a simple, screen-type reddening correction to the spectrum of Circinus and NGC~6240 is problematic since the observed Balmer decrement in both galaxies is very large (e.g. Fosbury and Wall 1979 for the case of NGC~6240). In addition, much of the \\Hb\\ flux in NGC~6240 is likely due to luminous star--forming regions in this galaxy. In both cases, and probably in many other narrow--line galaxies showing large Balmer decrements, the geometry is rather complex with several clouds along each line--of--sight. We may be looking into dusty environments for which a simple correction factor is inappropriate. This can invalidate the reddening-corrected \\Hb\\ fluxes used here. The theoretical I(Fe \\Ka)/I(\\Hb) is easily obtained from the spectral energy distribution, the column density and the iron abundance. For low ionization, small Balmer optical depth gas, the number of \\Hb\\ photons is a known fraction (about 0.12 for Case B recombination) of the Lyman continuum photon flux and the number of Fe \\Ka\\ photons is a known fraction of the ionizing E$>7.1$ keV flux. The case of interest for this study is gas with very large Lyman optical depth yet relatively small hard X-ray optical depth. Defining Q$_{7.1~keV}$ as the photon flux above 7.1 keV, and Q$_{13.6~eV}$ as the Lyman continuum photon flux, we get for this case \\begin{equation} {\\rm I(Fe \\Ka)/I(\\Hb)} \\simeq 1.5 \\times 10^3 \\exp(- \\tau_{\\rm 6.4 keV}) [\\frac{ {\\rm Q_{7.1~keV}}}{ {\\rm Q_{13.6~eV}}} ] [1-\\exp(-\\tau_{\\rm 7.1 keV})] \\,\\, , \\end{equation} where \\begin{equation} \\tau_{\\rm 7.1 keV} \\simeq 0.13 [\\frac{ {\\rm N_{col}} }{ 10^{23} } ] \\frac { ({\\rm n_{Fe}/n_H)} }{ 4 \\times 10^{-5} } \\,\\, , \\end{equation} is the 7.1 keV optical depth due to iron (assumed to be much smaller than unity), and $\\tau_{\\rm 6.4 keV}$ is the absorption optical depth due to all metals, at 6.4 keV (of the same order as $\\tau_{\\rm 7.1 keV}$ for solar ${\\rm n_{Fe}/n_H}$). Thus at small $\\tau_{\\rm 7.1 keV}$, the line ratio increases with the iron abundance. Major complications arise due to absorption of the E$>7.1$ keV photons by elements other than iron and by the non-negligible opacity at 6.4 keV, resulting in the destruction of emitted \\Ka\\ photons. This makes the Fe \\Ka\\ emissivity sensitive to the covering fraction since the \\Ka\\ photons emitted by one cloud can be absorbed by another cloud. Having in mind the NLR conditions, we assume in the following \\Cf=0.1. Fig.~8 shows a series of calculated I(Fe \\Ka)/I(\\Hb) for N$_{\\rm E} \\propto {\\rm E}^{-1.9}$ continuum, solar metallicity, \\nh=10$^4$\\cc\\ applicable to the NLR, and various \\aox. To enable a comparison with the expressions given above, we note that for those models with \\aox=1.3, ${\\rm Q_{0.1-10~keV}}/{\\rm Q_{7.1~keV}} = 130$ and ${\\rm Q_{13.6~eV}}/{\\rm Q_{0.1-10~keV}}=52$. The diagram shows that for \\Ux\\ $=10^{-3}$ and \\Ncol$>10^{21.7}$ \\cmii, \\Hb\\ is already emitted at maximum efficiency while the Fe \\Ka\\ flux is proportional to the column density. For \\Ncol$>10^{23.5}$ \\cmii, both lines are emitted at maximum efficiency and their ratio reflects the spectral energy distribution, ${\\rm n_{Fe}/n_H}$ and the destruction of the Fe \\Ka\\ photons. Inspection of Fig.~8 and Table 1 suggests that in four sources, NGC~1068, NGC~6240 Mrk~3 and Mkn~348, the I(Ka)/I(\\Hb) line ratio is below 0.2, which is consistent with solar ${\\rm n_{Fe}/n_H}$ for \\Ncol$< 10^{23.5}$ \\cmii\\ for \\aox=1.1. Assuming an iron overabundance by 2--3 reduces the required column to below 10$^{23}$ \\cmii\\ for the same \\aox. Furthermore, in three of the four cases the required column can be substantially smaller than the above mentioned upper limit. The column density is smaller if the UV bump is weaker than assumed and larger if \\aox\\ is larger than assumed. Thus in about half the sources the 6.4 keV iron line could originate in the NLR if the clouds in that region have column densities exceeding about 10$^{22.5}$ \\cmii. Circinus and NGC~4388 are different since I(Fe \\Ka)/I(\\Hb) in those sources is an order of magnitude larger than in the other three. As shown in the diagram, this is unlikely to be due to a much larger column density. Either the X-ray source is very bright compared with the UV source (very small \\aox) or else there is an additional, large covering fraction, \\Ka\\ producing component which is neutral, very thick and inefficient \\Hb\\ emitter. Current NLR models (see Ferguson \\etal\\ 1997 and references therein) do not make definite predictions regarding the size of the NLR clouds, since most observed narrow lines originate in the highly ionized, H{\\sc ii} part of the clouds. An obvious complication is if the NLR gas is dusty (see Netzer and Laor 1993 and references therein). For example, it is conceivable that the NLR clouds are the illuminated faces of dusty molecular clouds of significant column density. This would result in a reduced \\Hb\\ emissivity but the Fe \\Ka\\ line will hardly be affected. As already explained, the reddening correction for a dusty H{\\sc ii} region can differ substantially from the correction procedure applied here. We have examined the much larger sample in Turner \\etal\\ (1997a) to estimate the intrinsic I(Fe \\Ka)/I(\\Hb) ratio in type II AGN. Out of 17 sources with reliable Fe \\Ka\\ and \\Hb\\ measurements (including the ones in this paper), 7 show reddening corrected I(Fe \\Ka)/I(\\Hb)$<0.2$ which we consider consistent with origin in the NLR of these galaxies. The remaining sources show a larger ratio that requires \\Rf\\ in excess of what is expected from the NLR gas. The neutral Fe \\Ka\\ line in those is likely due to absorption by a larger column density, very neutral material that is either the walls of the central torus or large molecular clouds in the nucleus. Given the likely origin of the 6.4 keV line, we can now comment on the nature and location of the gas producing the high ionization iron lines. Assuming the same ${\\rm n_{Fe}/n_H}$ in both components, we can derive \\Rf(hot)/\\Rf(cold). This is found to be about 1 for NGC~1068 and about 0.1 for Mkn~3 and Circinus. The uncertainty is about a factor 2 since, as explained, the scattering efficiency differs by about this factor when comparing Compton-thin and Compton-thick gas. We further consider possible combinations of covering factor and note that for Compton-thin gas, \\Rf$\\propto$\\Ncol$\\times$\\Cf. This, combined with \\Ux(cold)/\\Ux(hot) (about 10$^{-3}$ with a large uncertainty, see Fig. 1 and the parameters used earlier for the NLR gas), enables us to estimate several likely combinations of these quantities. An interesting possibility involving the NLR idea, is that \\Ncol(hot)$\\sim 10^{-2}$\\Ncol(cold). This would imply \\Cf(hot)$\\simeq 10$\\Cf(cold), i.e. \\Cf(hot)$\\simeq$1. In this case, the hot and cold components coexist, spatially, and the large \\Cf(hot) does not allow a torus with a small opening angle. The typical NLR density is about $10^4$~\\cc, thus \\nh(hot)$\\sim 10 {\\rm cm^{-3} }$. The physical thickness of the hot gas in this case is of order 10-100 pc, i.e. of the same order of the NLR size. We note, however, that the two components are not in pressure equilibrium since nkT$_{\\rm e}$(cold)$\\simeq 10$nkT$_{\\rm e}$(hot). Another possibility is that \\Cf$\\sim 0.1$ in both components and \\Ncol(cold)$\\simeq 10$\\Ncol(hot). This does not allow a co-spatial existance of the two components. Finally, the 6.4 keV line may be from the thick torus walls. The efficiency in this case is very large and suggests that the fraction of this wall visible to us is extremely small. It also suggests that the hot gas completely fills the opening in the torus. Acknowledgments: It is a pleasure to acknowledge stimulating and useful discussions with our colleagues R. Mushotzky, K. Nandra, T. Yaqoob and T. Kallman. A very useful referee report helped us improve the presentation of this paper. This research is supported by the Universities Space Research Association (TJT, IMG) and by a special grant from the Israel Science Foundation (HN). \\newpage" }, "9803/astro-ph9803027_arXiv.txt": { "abstract": "New multi-epoch, mid-infrared (8-13\\,$\\micron$) spectrophotometric observations are presented for 30~late-type stars. The observations were collected over a four year period (1994-1997), permitting an investigation of the mid-infrared spectral shape as a function of the pulsation cycle (typically 1-2 years). The spectra of stars with little excess infrared emission and those with carbon-rich dust show the least spectral variability, while stars with evidence for dusty, oxygen-rich envelopes are most likely to show discernible variations in their spectral profile. Most significantly, a large fraction of variable stars with strong 9.7\\,$\\micron$ emission features show clear spectral profile changes which repeat from one cycle to the next. The significant sharpening of the silicate feature near maximum light can not be fully explained by heating and cooling of the circumstellar dust shell during the pulsational cycle, suggesting that the dust optical properties themselves must also be varying. In addition, the appearance of a narrow emission feature near the silicate peak for a few stars may require the production of especially ``pure'' silicate dust near maximum light. The general narrowing of the silicate feature observed may reflect the evolution of the pre-existing dirty grains whose surface impurities have been evaporated off when the grain temperature rises preceding maximum light. An improved theory of dust formation which can explain the observed changes in the grain properties around a single, pulsating star may lead to a definitive explanation for the diversity of silicate emission profiles observed amongst oxygen-rich, late-type stars. ", "introduction": "The mid-infrared (8-13\\,$\\micron$) spectra of late-type stars have been measured by many observers since the development of infrared detectors. These red giants and supergiants are often surrounded by dusty envelopes which absorb stellar radiation and re-radiate the energy in the near- and mid-infrared. The infrared spectra can be classified based on the chemical content of the circumstellar environment (oxygen- or carbon-rich) and on the optical thickness of the dusty envelope (e.g., Merrill \\& Stein 1976a,b). Oxygen-rich circumstellar environments often produce spectra evincing a feature near 9.7\\,$\\micron$ resulting from the presence of silicate dust (Woolf \\& Ney 1969). This feature appears in emission for optically thin envelopes or in absorption when large enough optical depths are encountered. The emission spectra of dust surrounding carbon stars are nearly featureless, although often containing an 11.3\\,$\\micron$ feature attributed to SiC. In addition, many of these red giants and supergiants are classified as long-period variables, pulsating with a typical period of 1-2 years. The homogeneous set of survey measurements by the Infra-Red Astronomical Satellite (IRAS) in the mid-1980s allowed observers to classify silicate emission features based on various schemes (IRAS Science Team 1986; Little-Marenin \\& Little 1988, 1990; Goebel et al. 1989; Sloan \\& Price 1995). The different shapes of the feature have been interpreted largely as due to differences in the chemical make-up of oxygen-rich dust. Unfortunately the IRAS program did not conscientiously include observations of the mid-infrared spectra of long-period variable stars at different phases of their luminosity cycles, and there are only a few cases where such data have been retrieved from the IRAS Low Resolution Spectrometer (LRS) database. These observations have suggested silicate feature strength variations as a function of pulsational cycle, but have been hampered by limited temporal coverage (Little-Marenin, Stencel, \\& Staley 1996). More recent results by Creech-Eakman et al. (1997) point towards evidence for variations in the silicate feature as a function of pulsational phase, but the comparison spectra were taken nearly a decade apart. Hence, the simple observational question of whether the mid-infrared spectra of LPVs change shape through the pulsational cycle has been left without a decisive answer. A campaign of observations taken from 1994 to 1997 was designed to monitor the mid-infrared spectrum of nearly 30 late-type stars. The observations, sampling the spectrum of most stars multiple times within a pulsational cycle, used the same instrument and observing technique. The homogeneity of this data set is important for allowing reliable spectral comparisons, avoiding complicating issues such as different apertures and calibration methods. This paper presents the full data set collected thus far and discusses the spectral variability (or lack of variability) of our sample stars. ", "conclusions": "The mid-infrared spectra of 30~late-type stars have been monitored in order to detect changes occurring on the pulsational time scale (typically 1-2~years) of long period variables (LPVs). Stars which exhibited little or no bolometric variability (i.e. non-LPVs) generally showed no change in their spectral shape in the range 8-13\\,$\\micron$. Furthermore, most stars with no strong 9.7\\,$\\micron$ silicate feature, including carbon stars and oxygen-rich miras with broad, weak silicate features, showed no spectral shape change. However, a few such stars in this category displayed either enhanced variability as a function of wavelength (IRC~+10216 and R~Leo) or a detectable change in the spectral slope correlated with pulsational phase (R~Cnc and W~Aql). The former effect has no clear explanation, while the latter effect can be explained by changes in the dust shell temperature ($\\Delta T_{\\rm{dust}}\\simle 200$\\,K). The most significant result presented here is that nearly all of the observed sources with clear 9.7\\,$\\micron$ silicate features and definite bolometric variability showed strong evidence for changes in the silicate emission strength and spectral profile which are correlated with pulsational phase. We conclude that silicate emission variation is a general property of long-period variables with optically thin silicate features. The sharpening of the silicate feature near maximum light and its subsequent broadening can be explained by the heating and cooling of the dust envelope coupled with changing optical constants for the dust grains. The appearance of a spectrally narrow emission feature near the silicate peak of a few stars strongly indicates the existence of ``pure'' silicate dust grains near maximum light. We hypothesize that the general narrowing of the dust emission spectra may arise from pre-existing dirty grains whose surface impurities have been evaporated off or whose amorphous molecular configuration has crystallized during the dust re-heating following minimum luminosity. The solid-state resonance would naturally broaden after maximum light as impurities re-adsorb onto the cooled grain surface. Such speculation awaits more detailed laboratory measurements of astrophysically-relevant grain types. The observations presented here remind us that the dust formation process is still only partially understood. Indeed, uncertainties in the optical constants for circumstellar dust are a primary obstacle in creating self-consistent multi-wavelength radiative transfer models incorporating interferometric observations of dusty objects (e.g., Monnier et al. 1997). The changes observed in silicate optical properties as a function of pulsational phase are not predicted by any present dust formation theories, and more careful consideration is required of the effects of photospheric shocks propagating into the dust formation zone and of the changing temperature and density structure due to the pulsation. Such models may then not only explain the changing optical properties of the dust around a single, pulsating object, but may also explain why different stars possess silicate emission with distinctly different spectral profiles." }, "9803/astro-ph9803094_arXiv.txt": { "abstract": "The detection of rapid variability on a time\\-scale of hours in radio-quiet quasars (RQQSOs) could be a powerful discriminator between starburst, accretion disc and relativistic jet models of these sources. This paper contains an account of a dedicated search for rapid optical variability in RQQSOs. The technique used differential photometry between the RQQSO and stars in the same field of view of the CCD. The 23 RQQSOs that were observed all have high luminosities ($-271$. The total amount of observation time was about 60 hours and these observations are part of an ongoing programme, started in September 1990, to search for rapid variability in RQQSOs. No evidence for short-term variability greater than about 0.1 magnitudes was detected in any of the 23 sources, however long-term variability was recorded for the radio-quiet quasar \\object{PG 2112$+$059}. The finding charts are included here because they identify the RQQSO and the reference stars used in the photometry, and hence are available for use by other observers. The unusual properties of two RQQSOs that were not included in our source list are noted. X-ray results reveal that \\object{PG 1416$-$129} is variable on a timescale of days and that the remarkable source \\object{IRAS 13349$+$2438} varied by a factor of two on a timescale of a few hours. The latter source displayed blazar type behaviour in X-rays and implies that relativistic beaming may occur in at least some RQQSOs. Radio results also indicate the presence of jets in at least some RQQSOs. ", "introduction": "There is a general consensus that quasars belong to two different radio populations, radio-quiet quasars (RQQSOs) and radio-loud quasars. $R$ is usually defined as the ratio of the radio (6~\\mbox{cm}) to the optical (440~\\mbox{nm}) flux densities and the radio-quiet quasars have a value of $R<10$, while the radio-loud quasars have $R>10$ (Kellermann et al. 1989). It is found that $\\sim10~\\mbox{\\%}$ of quasars are in the radio-loud category. An additional distinction between active galactic nuclei (AGN) with strong and weak radio sources comes from the observation that radio loud objects essentially all occur in elliptical galaxies and RQQSOs appear to reside in galaxies that are dominated by exponential disks. However the RQQSOs that occur in elliptical host galaxies are in general more luminous than those that reside in disks (Taylor et al. 1996). Little is known about the short-term variability of ra\\-dio-quiet quasars, because few studies have been carried out (Gopal-Krishna et al. 1993 and 1995; Jang \\& Miller 1995; Sagar et al. 1996). In contrast blazars display rapid variability in the wavelength range from radio to gamma rays. The blazar class encompasses both optically-violent\\-ly-variable (OVV) quasars and BL Lac objects and about one quarter of all radio-loud quasars are also in the blazar category (Webb et al. 1988; Pica et al. 1988). There are many theoretical models which endeavour to explain the large and rapid variability exhibited by blazars and these are usually divided into extrinsic and intrinsic categories. One extrinsic mechanism is microlensing of emission knots in a relativistic jet when they pass behind planets in an intervening galaxy (McBreen \\& Metcalfe 1987; Gopal-Krishna \\& Subramanian 1991). The rapid variability from superluminal-microlensing may be \\linebreak responsible for the variability observed in \\object{AO 0235$+$164} (Rabbette et el. 1996) and \\object{PKS 0537$-$441} (Romero et al. 1995). One family of intrinsic models is based on a rotating supermassive black hole which accretes matter from a surrounding accretion disc and ejects two oppositely directed jets. The shocked-jet model involves shocks which move with relativistic speeds along the jet (Qian et al. 1991; Marscher, 1980). It is believed the shock propagates along the line of sight, through inhomogeneous, small-scale structures distributed along the jet. These inhomogeneous structures are illuminated, or excited, by the moving relativistic shock, through the amplification of the magnetic field and the acceleration of electrons which causes the variability in polarization and in flux density that are observed over a wide range of frequencies (Hughes et al. 1986). Another family of intrinsic models invokes numerous flares or hotspots in the accretion disk and the corona that is believed to surround the central engine (Wiita et al. 1992; Mangalam \\& Wiita 1993) and indeed a similar model has been proposed to explain X-ray variations in blazars (Abramowicz et al. 1991). The fact that RQQSOs generally lie on the far-infrared versus radio correlation (Sopp \\& Alexander 1991) suggest that star formation plays an important role in their radio emission. It has been suggested by Terlevich et al. (1992) that the low values of $R$ in RQQSOs can be explained without jets or accretion discs, by postulating a circumnuclear starburst within a dense, high-metallicity nuclear environment. In this model the optical/UV and bolometric luminosity arises from young stars; the variability comes from cooling instabilities in the shell of compact supernova remnants and supernova flashes. Variability on intranight timescales is however difficult to explain with this model because of the short timescales involved. Furthermore radio-quiet and radio-loud quasars have very different radio power outputs but have similar spectral shapes in the radio region and suggest that a significant fraction of the RQQSOs may be capable of producing powerful radio emission (Barvainis et al. 1996). Kellermann et al. (1994) found possible radio extensions up to about 300~\\mbox{kpc} in a few RQQSOs and assert that for at least these few cases, the emission is too large to be starburst related (Stein 1996). Recently, some evidence suggesting rapid optical variability in the RQQSOs \\object{PG 0946$+$301} and \\object{PG 1444$+$407} was reported by Sagar et al. (1996). They also reported long-term variability for four RQQSOs. Jang \\& Miller (1995) reported intranight variability for one RQQSO out of a sample of nine sources. Brinkmann et al. (1996) obtained ASCA observations of the radio-quiet, infrared \\linebreak quasar \\object{IRAS 13349$+$2438} and detected substantial X-ray variability on a timescale of only a few hours. The results of the photometric observations of a sample of mainly high luminosity and high redshift RQQSOs are presented. The observations and data reduction are given in Sect.~2. The results including tables listing the differential photometry and some light curves are presented in Sect.~3. The discussion and conclusions are given in Sects.~4 and 5. Sect.~4 also includes a discussion on two remarkable RQQSOs, \\object{PG 1416$-$129} and \\object{IRAS 13349$+$2438}. CCD images of the fields containing the radio-quiet quas\\-ars and reference stars used in the differential photometry are also included. A value of $\\mathrm{H}_\\mathrm{0}=50~\\mbox{km s}^{-1}~\\mbox{Mpc}^{-1}$ and $\\mathrm{q}_\\mathrm{0}=0.5$ has been adopted. ", "conclusions": "A long-term survey of a sample of high luminosity ($-27$1) likely indicate optically thin gas created by the high temperatures caused by star forming regions in the nucleus of this starburst galaxy. ", "introduction": "The circumnuclear regions of galaxies are very often the setting for starbursts and other extraordinary events. Observations of the gas kinematics and distribution indicate that bars, resonances, gas inflow, and tidal shear play important roles in the formation and evolution of nuclear starbursts (\\eg\\ Handa \\etal\\ 1990; Kenney \\etal\\ 1992). Previous observations show that the molecular gas in the central regions of barred spiral galaxies often does not extend all the way into the centre of the nucleus. It accumulates some distance away from the centre, giving the emission a double peaked appearance with each peak occurring where the bar meets the nucleus (Kenney \\etal\\ 1992; Ishizuki \\etal\\ 1990). Dynamical models indicate that in the presence of a barred potential, gas will flow inward along the bar and slow its descent temporarily near inner Lindblad resonances (ILR, \\eg\\ Combes 1988; Shlosman, Frank \\& Begelman 1989). At these locations the gas may accumulate into larger complexes of molecular clouds. In addition to complex dynamics, there is most likely a profusion of complicated photo-chemistry occurring within the nuclei of starburst galaxies. Interstellar clouds are believed to consist of smaller high density dark cores interspersed throughout a larger region of lower density. These high density regions are self-shielded from ultraviolet (UV) radiation that tends to dissociate molecular gas. The result is the population of the diffuse region by hydrogen atoms (\\ion{H}{1}), atomic and ionized carbon (C, C$^+$), and many other atomic and ionized species (\\eg\\ Morton \\etal\\ 1973), while the dark cores can contain molecular species such as H$_2$ and CO. It is believed that atomic carbon can exist only in a small energy window, outside of which it will either be ionized or combined with oxygen to form CO. Inside the dense cores, most of the carbon combines to form CO, while outside the cores, the UV radiation acts to ionize atomic carbon. It is therefore expected that \\ion{C}{1} is the dominant species near the edges of dense, self-shielding cloud cores. This simple model fails to explain the extended \\ci\\ emission observed in molecular clouds in our own Galaxy (Plume, Jaffe, \\& Keene 1994; Keene \\etal\\ 1985). One possible explanation is found in the clumpy structure of molecular clouds (\\eg\\ Stutzki \\& G\\\"usten 1990). This clumpiness would allow UV radiation to penetrate much deeper into the cloud allowing atomic carbon to exist at depths greater than would be allowed by a simple spherical model of molecular clouds (\\eg\\ Boisse 1990). Many alternative explanations have been proposed to explain the extended \\ci\\ emission. These ideas range from complicated chemical processes involving H$^+$ (Leung, Herbst, \\& Heubner 1984) to simpler ideas such as a C/O ratio greater than one (\\eg\\ Keene \\etal\\ 1985). This paper presents \\ci\\ and CO maps of the barred spiral galaxy M83. Its low inclination angle ($i$ = 24\\arcdeg, Comte 1981) and close proximity ($D$ = 4.7 Mpc, Tully 1988) make it one of the best locations for studying the response of gas to a barred potential. It is believed to be undergoing a nuclear starburst (\\eg\\ Talbot \\etal\\ 1979), which may have been triggered by molecular gas inflow along the bar potential. This nuclear starburst would produce higher temperatures which should readily excite the higher $J$-transitions in the CO gas. Also, strong UV flux has been detected in the nucleus (Bohlin \\etal\\ 1983), which would help dissociate the CO into atomic carbon. By studying the CO and \\ci\\ data, we can understand better the dynamics of the gas and its role in fueling the nuclear starburst and also learn about the conditions conducive to the formation of atomic carbon. ", "conclusions": "} This paper presents \\ci, \\twcoft, and \\twcott\\ maps of the nucleus of the barred spiral galaxy M83 taken at the JCMT. The main results are summarized below. \\begin{enumerate} \\item{We observe a double peaked structure in the molecular emission consistent with gas inflow along the bar collecting at the inner Lindblad resonance. The \\twcoft\\ emission suggests that some of the molecular gas has made it into the nucleus and is being heated by and possibly fueling the nuclear starburst. This result indicates that nuclear starbursts may occur even in galaxies which exhibit a double peaked emission structure, in contrast to the findings of Kenney \\etal\\ (1992).} \\item{We observe different morphologies in the \\twcoft\\ channel maps than in the \\twcott\\ and \\twcooz\\ channel maps. These data suggest that \\twcooz\\ emission may not always be a good tracer of molecular gas in starburst galaxies, as the CO may be heated sufficiently to produce little emission in the \\joz\\ line. Thus, discretion should be applied in the interpretation of \\twcooz\\ emission as a tracer of molecular gas in starburst regions.} \\item{The observations also suggest that the double peaked emission may be the result of a molecular ring out of the plane of the galaxy oriented nearly perpendicular to the main disk. This torus of cooler gas would need to contain a disk of hotter gas that fills its central void in order to explain the observed morphology.} \\item{The \\ci\\ line strength indicates carbon column densities of $1 \\times 10^{18}$ cm$^{-2}$ while CO emission indicates CO column densities of $\\sim 3 \\times 10^{18}$ cm$^{-2}$ and H$_2$ column densities of $3 \\times 10^{22}$ cm$^{-2}$ at the peak of the emission. The $N$(C)/$N$(CO) ratio at this location is 0.33 $\\pm$ 0.10 which is similar to those found in other starburst galaxies.} \\item{The twelve \\ci/\\twcoft\\ line ratios in the inner 24$''\\times 24''$ are uniform at the 2$\\sigma$ level and have an average value of 0.25 $\\pm$ 0.03, similar to those of other starburst galaxies. The uniformity of the line ratios suggests that both the high-excitation CO emission and atomic carbon form in photodissociation regions in the starburst nucleus.} \\item{The \\twcoft/\\twcott\\ integrated intensity line ratios vary substantially over the central region of M83 with the highest ratios occurring towards the edges of the emission peaks. The \\twcoft/\\jtt\\ line ratios seem to be enhanced along an arc of active star forming regions. The high line ratios ($>$1) indicate that the higher $J$ transitions of CO are optically thin and are likely the result of the high temperatures and/or densities associated with star formation. } \\end{enumerate}" }, "9803/astro-ph9803006_arXiv.txt": { "abstract": "Using the IRAM interferometer we have mapped at high resolution ($2\\farcs2 \\times 1\\farcs2$) the $^{12}$CO(1--0) emission in the nucleus of the doubled barred SABbc spiral M~100. Our synthesized map includes the zero spacing flux of the single--dish 30m map (Sempere \\& Garc\\'\\i a--Burillo, 1997, {\\bf paper I}). Molecular gas is distributed in a two spiral arm structure starting from the end points of the nuclear bar ($r=600$ pc) up to $r=1.2$ kpc, and a central source ($r\\sim$100 pc). The kinematics of the gas indicates the existence of a steep rotation curve (v$_{rot}$=180 km\\, s$^{-1}$ at $r\\sim 100$ pc) and strong streaming motions characteristic of a trailing spiral wave inside corotation. Interpretation of the CO observations and their relation with stellar and gaseous tracers (K, optical, H$\\alpha$, H\\,I and radiocontinuum maps) are made in the light of a numerical model of the clouds hydrodynamics. Gas flow simulations analyse the gas response to a gravitational potential derived from the K-band plate, including the two nested bars. We develop two families of models: first, a single pattern speed solution shared by the outer bar+spiral and by the nuclear bar, and secondly, a two independent bars solution, where the nuclear bar is dynamically decoupled and rotates faster than the primary bar. We found the best fit solution consisting of a fast pattern ($\\Omega_f$=160 kms$^{-1}$kpc$^{-1}$) for the nuclear bar (with corotation at R$^{F}_{COR}$=1.2 kpc) decoupled from the slow pattern of the outer bar+spiral ($\\Omega_f$=23 kms$^{-1}$kpc$^{-1}$) (with corotation at R$^{S}_{COR}$=8-9 kpc). As required by non-linear coupling of spirals (Tagger et al 1987), the corotation of the fast pattern falls in the ILR region of the slow pattern, allowing an efficient transfer of molecular gas towards the nuclear region. Solutions based on a single pattern hypothesis for the whole disk cannot fit the observed molecular gas response and fail to account for the relation between other stellar and gaseous tracers. In the two-bar solution, the gas morphology and kinematics are strongly varying in the rotating frame of the slow large-scale bar, and fit the data periodically during a short fraction (about 20\\%) of the relative nuclear bar period of 46 Myr. ", "introduction": "The advent of high-sensitivity near-infrared imaging of galaxies has shown that a significant percentage of barred spirals host secondary bars in their nuclei. There could be two interpretations of the {\\it bars within bars} phenomenon, according to the relative pattern speeds of the two bars (Friedli and Martinet, 1993; Friedli and Benz, 1993 and 1995; Combes, 1994). The two patterns could be corotating if they are about parallel or perpendicular to each other. If the secondary inner bar is strongly misaligned with the primary outer bar, the two bars are likely to have distinct wave pattern speeds, as shown by numerical simulations. The decoupling of an inner faster pattern appears in self-consistent simulations with gas and stars thanks to the role of the dissipative component: as a result of gas inflow, under the action of the bar gravitational torques, mass accumulates onto the x$_2$ types of orbits which weakens the primary bar. The rotation period becomes much shorter in the nuclear regions due to mass concentration, which leads to the decoupling of a fast-rotating bar. Eventually, the nuclear bar destroys itself or it destroys the primary bar, modifying the overall disk potential. Evolution can then occur in much less than a Hubble time, and galaxies change their morphological type along the Hubble sequence. They change from barred to un-barred, and also they concentrate mass in the process, evolving slowly from late-types to early types. The observation and modelling of {\\it real} barred galaxies offers the opportunity to test theory predictions on galaxy evolution and it seems a necessary complement to numerical simulations of {\\it model} galaxies. The present work is intended to bring a combined observational and modelling effort in the study of the nearby barred spiral M100 (NGC4321). In this galaxy, classified as SABbc by de Vaucouleurs et al (1991), the hypothesis of a single mode common to the whole disk is dubious, both from observational and theoretical evidences. \\bigskip On the observational side, the nuclear region of M100 (up to r=3kpc) has been so far the subject of numerous studies. The pioneering work of Arsenault and collaborators (1988, 1989, 1990) established a connection between the ring-like H$\\alpha$ morphology of the nucleus and the existence of Inner Linblad Resonances. Further steps in sensitivity made appear, first, a four-armed structure (Cepa et al, 1990) and recently a fragmented two spiral arm structure in H$\\alpha$ (Knapen et al, 1996). The 6cm radio-continuum VLA map of Weiler et al (1981) shows also a two arm spiral pattern. Near infrared images of the nucleus (Pierce 1986, Shaw et al 1995, Knapen et al 1995 (hereafter {\\bf K95}), Rauscher 1995) show the existence of either a secondary nuclear bar in K (nearly parallel to the main bar) together with a leading spiral structure, or an inner oval in the I band (with principal axes misaligned with respect to the outer bar). The synthesis aperture $^{12}$CO(1--0) maps (Canzian, 1992; Rand, 1995; Sakamoto et al 1995) indicate the existence of a two-arm molecular spiral structure connected to the K nuclear bar end points. The IRAM 30m map of {\\bf paper I} shows a strong concentration of CO emission towards the nuclear disk {\\bf ND}, a component clearly distinguishable from the main bar. A steep rotation curve gradient, unresolved by the 30m beam (12\\arcsec\\, in the 2--1 line), indicates a high mass concentration in the {\\bf ND}. \\bigskip On the modelling side, Garc\\'\\i a-Burillo et al (1994) (hereafter called {\\bf GB94}) and Sempere et al (1995) (hereafter {\\bf S95}) made numerical simulations of the cloud hydrodynamics to study the evolution of the molecular gas disk under the action of a realistic spiral+barred potential derived from a red band plate. The authors assume the whole disk to be fitted by a single well defined wave pattern characterized by $\\Omega_p$, shared by the primary bar and the spiral arms. However they lacked first, of the necessary spatial resolution and secondly, of a fair potential tracer to analyse the gas response in the inner 500 pc. {\\bf K95} have made numerical simulations of the stellar and gas dynamics in M100, using a {\\it model} potential which departs markedly from the real mass distribution. Although they favour a one bar mode scenario their model fails to reproduce the molecular gas distribution observed by the interferometer. \\bigskip We present here a combined single-dish and interferometer data set fulfilling both high resolution (2.2\\arcsec$\\times$1.2\\arcsec) and sensitivity requirements. Contrary to the synthesis aperture maps so far published, we recover entirely the zero-spacing flux of the {\\bf ND}. The comparison between the different gaseous and stellar tracers of the {\\bf ND} is reexamined in this work. Particular attention is paid to the bias introduced by extinction in optical and even near-infrared images of the nucleus, and what might be the implications on the interpretation of the data. Observations are confronted to the result of new numerical simulations of the clouds hydrodynamics, based on a mass distribution directly derived from the infrared luminosity image of M100. We will focus on the feasibility of two independent patterns in the disk and how this scenario accounts better for the observations. ", "conclusions": "Simulations of the H$_2$ cloud hydrodynamics in the double barred system M100 have shown that the ensemble of observations (optical, infrared, HI and CO maps) are best explained by a two {\\it independent bars} scenario. The primary stellar bar (of 4.5 kpc radius) and the outer spiral structure share a common pattern speed of $\\Omega_s$=23\\,kms$^{-1}$kpc$^{-1}$ which places corotation at R$_{COR}^S$=8-9 kpc, i.e. beyond the bar end-points though well inside the optical disk. Although the nuclear stellar bar is mostly aligned with the primary bar (within 20\\deg) it has been shown to lead a fast pattern rotating at $\\Omega_f$=160\\,kms$^{-1}$kpc$^{-1}$, having corotation at R$_{COR}^{F}$=1.2 kpc radius. Both modes are dynamically decoupled and they show overlapping of their major resonances: corotation of the fast mode falls well within the ILR region of the slow mode. The present model explains the efficient gas transport towards the nucleus, suggested by the interferometer observations, as a consequence of secular evolution driven by the stellar bar. Molecular gas crosses the ILR region of the slow pattern, spiraling inwards and forming a trailing spiral structure and an ultracompact source encircled by the ILR of the fast pattern (R$_{iILR}^{F}$=2.5$\\arcsec$). Alternative solutions are unable to account for the CO observations. In particular, in the slow pattern solution gas is stopped at the ILR barrier and forms a nuclear ring outside the {\\bf ND} extent. No central gas condensation is formed either. The fast pattern solution proposed by K95 ($\\Omega_p$=70\\,kms$^{-1}$kpc$^{-1}$) worsens the fit for the outer bar+spiral structure found by {\\bf GB94}. In addition, two independent methods based on the morphology of the residual velocity field for the gas ({\\bf S95}) and the identification of spurs in optical pictures (e.g. Elmegreen et al 1992) confirm the value of R$_{COR}^S$ reported above. We conclude that the gas response derived from the CO interferometer map, and the relation between the different stellar (K image) and gaseous tracers of the {\\bf ND} (H$\\alpha$) are best explained by the two pattern model. In particular, it explains the high CO concentration in the central part. This gas concentration could be eventually the cause of the nuclear bar destruction in this fastly evolving galaxy (see Norman et al 1996). {\\it Acknowledgements}. This work has been partially supported by the Spanish CICYT under grant number PB96-0104. We thank J. Knapen for providing us with the H\\,I, H$\\alpha$ and K-band images used in this paper." }, "9803/astro-ph9803140_arXiv.txt": { "abstract": "The Westerbork Northern Sky Survey (WENSS) has been used to select a sample of Gigahertz Peaked Spectrum (GPS) radio sources at flux densities one to two orders of magnitude lower than bright GPS sources investigated in earlier studies. Sources with inverted spectra at frequencies above $325$ MHz have been observed with the WSRT\\footnote{The Westerbork Synthesis Radio telescope (WSRT) is operated by the Netherlands Foundation for Research in Astronomy with financial support from the Netherlands Organisation for Scientific Research (NWO).} at 1.4 and 5 GHz and with the VLA\\footnote{The Very Large Array (VLA) is operated by the U.S. National Radio Astronomy Observatory which is operated by the Associated Universities, Inc. under cooperative agreement with the National Science Foundation.} at 8.6 and 15 GHz to select genuine GPS sources. This has resulted in a sample of 47 GPS sources with peak frequencies ranging from $\\sim$500 MHz to $>$15 GHz, and peak flux densities ranging from $\\sim$40 to $\\sim$900 mJy. Counts of GPS sources in our sample as a function of flux density have been compared with counts of large scale sources from WENSS scaled to 2 GHz, the typical peak frequency of our GPS sources. The counts can be made similar if the number of large scale sources at 2 GHz is divided by 250, and their flux densities increase by a factor of 10. On the scenario that all GPS sources evolve into large scale radio sources, these results show that the lifetime of a typical GPS source is $\\sim 250$ times shorter than a typical large scale radio source, and that the source luminosity must decrease by a factor of $\\sim 10$ in evolving from GPS to large scale radio source. However, we note that the redshift distributions of GPS and large scale radio sources are different and that this hampers a direct and straightforward interpretation of the source counts. Further modeling of radio source evolution combined with cosmological evolution of the radio luminosity function for large sources is required. ", "introduction": "Gigahertz Peaked Spectrum (GPS) radio sources are a class of extragalactic radio source characterized by a spectral peak near 1 Gigahertz in frequency (e.g. Spoelstra et al. 1985) The spectral peak in these compact luminous objects is believed to be due to synchrotron self absorption caused by the high density of the synchrotron emitting electrons in the radio source. GPS sources are interesting objects, both as Active Galactic Nuclei (AGN) and as cosmological probes. It has been suggested that they are young radio sources ($<10^4$ yr) which evolve into large radio sources (Fanti et al. 1995, Readhead et al. 1996, O'Dea and Baum 1997), and studying them would then provide us with important information on the early stages of radio source evolution. Alternatively GPS sources may be compact because a particularly dense environment prevents them from growing larger (e.g. O'Dea et al. 1991). Important information about the nature of GPS radio sources comes from the properties of their optical counterparts. The galaxies appear to be a homogeneous class of giant ellipticals with old stellar populations (Snellen et al. 1996a, 1996b, O'Dea et al. 1996) and are thus useful probes of galaxy evolution with little or no contamination from the active nucleus in the optical. GPS quasars have a different redshift distribution to their galaxy counterparts ($21$ Jy (Fanti et al. 1990, O'Dea et al. 1991, Stanghellini et al. 1996, de Vries et al. 1997). We are carrying out investigations of GPS sources at fainter flux density levels, in order to compare their properties with their radio bright counterparts. This enables us to investigate the properties of GPS sources as a function of radio luminosity, redshift, and rest frame peak frequency. The selection of a sample at intermediate flux densities was described in Snellen et al. (1995a). This paper describes and discusses the selection of an even fainter sample from the Westerbork Northern Sky Survey (WENSS, Rengelink et al. 1997). \\section {Selection of GPS Sources} \\subsection{The Westerbork Northern Sky Survey} The Westerbork Northern Sky Survey (WENSS) is being carried out at 325 and 609 MHz (92 and 49 cm) with the Westerbork Synthesis Radio Telescope (WSRT). At 325 MHz, WENSS covers the complete sky north of $30^\\circ$ to a limiting flux density of approximately 18 mJy ($5 \\sigma$). At 609 MHz, about a quarter of this area, concentrated at high galactic latitudes, has been surveyed to a limiting flux density of approximately 15 mJy ($5 \\sigma$). The systematic errors in flux density in WENSS were found to be $\\sim 5\\%$ (Rengelink et al. 1997). The survey was conducted in mosaicing mode which is very efficient in terms of observing time. In this mode, the telescope cycles through 80 evenly spaced field centres, during each of a number of $12^h$ syntheses with different spacings of array elements. The visibilities are sufficiently well sampled for all 80 fields that it is possible to reconstruct the brightness distribution in an area of the sky, $\\sim$100 square degrees, which is many times larger than the primary beam of the WSRT. Individual fields are referred to as {\\it mosaics}, and have a resolution (FWHM of the restoring beam) of $54'' \\times 54''$ cosec $\\delta$ at 325 MHz and $28''\\times 28''$ cosec $\\delta$ at 609 MHz. From the combined mosaics, maps are made with a uniform sensitivity and regular size, which are called {\\it frames}. The 325 MHz frames are $6^\\circ \\times 6^\\circ $ in size and positioned on a regular $5^\\circ \\times 5^\\circ $ grid over the sky, which coincides with the position grid of the new Palomar Observatory Sky Survey (POSS II, Reid et al. 1991) plates. A detailed description of WENSS is given by Rengelink et al. (1997) \\subsection{Selection of a Sample of Candidate GPS Sources.} A deep low frequency radio survey such as WENSS is crucial for selecting a sample of faint GPS sources. It is the inverted spectrum at low frequencies which distinguishes them from other types of radio sources. Figure \\ref{surveys} shows the major large-sky radio surveys in the northern sky with theoretical spectra of homogeneous synchrotron self absorbed radio sources (eg. Moffet 1975) superimposed, which have spectral peak frequencies of 1 GHz. Samples of GPS sources can be constructed using WENSS flux density measurements in the optically thick part of their spectra which are ten times fainter than samples selected using the Texas Survey (Douglas 1996). \\begin{figure}[!t] \\centerline{ \\psfig{figure=figure1.ps,width=8cm}} \\caption{\\label{surveys} Overview of the major radio surveys in the northern sky: the Greenbank Surveys (Condon and Broderick 1985, Gregory and Condon 1991), the Texas Survey (Douglas et al. 1996), and the Cambridge 3C, 4C, and 6C surveys. The curves represent the spectra of a homogeneous synchrotron self absorbed radio source, with a peak frequency of 1 GHz and peak flux density of 300 mJy (lower curve) and 3000 mJy (upper curve). Samples of GPS sources can be constructed using WENSS flux density measurements in the optically thick part of their spectra which are more than an order of magnitude fainter than samples selected using the Texas Survey.} \\end{figure} When we selected our sample, only a small part of the WENSS region had been observed and the data reduced to the point of providing source lists. The 325 MHz WENSS data used to select GPS sources are from two regions of the sky; one at $15^{\\rm h} < \\alpha < 20^{\\rm h}$ and $58^\\circ< \\delta < 75^\\circ$, which is called the {\\it mini-survey} region (Rengelink et al. 1997), and the other at $4^{\\rm h}00^{\\rm m} < \\alpha < 8^{\\rm h}30^{\\rm m}$ and $58^\\circ< \\delta < 75^\\circ$, where $\\alpha$ is right ascension and $\\delta$ is declination. These were the first two regions observed, reduced and analysed for WENSS. The mini-survey region, which is roughly centered on the North Ecliptic Pole, was chosen as the first area for analysis because it coincides with the NEP-VLA survey at 1.5 GHz (Kollgaard et al. 1994), the deep 7C North Ecliptic Cap survey (Lacy et al. 1995, Visser et al. 1995), the deepest part of the ROSAT All Sky survey (Bower et al., 1996) and the IRAS survey (Hacking and Houck, 1987). The high declination of the two regions is very convenient for VLBI experiments, because their locations are circumpolar for almost all the major EVN and VLBA radio telescopes. At the time of selection WENSS 609 MHz data was available for only about one third of the mini-survey region. The regions for which both 325 and 609 MHz source lists were available cover 119 square degrees of the sky. The regions for which only 325 MHz data were available cover 216 square degrees in the mini-survey region and 306 square degrees in the other region. These source lists were correlated with those from the Greenbank 5 GHz (6 cm) survey (Gregory and Condon 1991, Gregory et al. 1996), which has a limiting flux density of 25 mJy ($5 \\sigma$). For the faintest sources the new Greenbank source list (Gregory et al. 1996) was used, which is based on more data. Candidate GPS sources were selected on the basis of a positive spectral index $\\alpha$ between 325 MHz and 5 GHz, where the spectral index is defined by $S \\sim \\nu^{\\alpha}$. If 609 MHz data was also available, an ``inverted'' spectrum between 325 MHz and 609 MHz was used as the selection criterion. This in fact increased the sensitivity of the selection process to GPS sources with low peak frequencies ($< 1 $ GHz). Note that in general for a GPS source, the 325-609 MHz spectral index will be more positive than the 325-5000 MHz spectral index for a spectral peak in the 1 GHz range. Hence, using the 325-609 MHz selection criterion will not miss any GPS sources which would have been found using the 325-5000 MHz selection criterion, it will only add extra sources with lower peak frequencies. In total, 117 inverted spectrum sources were selected; 37 using the 325-609 MHz selection and 82 using the 325-5000 MHz selection. They are listed in table \\ref{canGPS}. Columns 1, and 2 give the name, right ascension and declination (B1950) (obtained from the VLA observations), columns 3, 4 and 5 the 325 MHz, 609 MHz and 5 GHz flux densities, and columns 6 and 7 give the 325-609 MHz and 325-5000 MHz spectral indices. The uncertainties in the 325-5000 MHz spectral indices range from 0.03 to 0.05 (for the faintest objects), and the uncertainties in the 325-609 MHz spectral index range from 0.10 to 0.40. \\subsection{Additional Radio Observations.} An apparently inverted or peaked spectrum could be caused by variability at any or all of the selection frequencies, due to the fact that the 325, 609 and 5000 MHz surveys were observed at different epochs. To select the genuine GPS sources, additional quasi-simultaneous observations at other frequencies are required to eliminate flat spectrum, variable radio sources. The 5 GHz Greenbank survey was made in 1987, while the 325 MHz and 609 MHz data were taken in 1993. Furthermore, high frequency data is needed to confirm their turnover, and measure the (steep) spectrum in the optically thin part of their spectra. Therefore VLA observations were taken at 8.4 and 15 GHz, and WSRT observations at 1.4 and 5 GHz. Later, after the selection process, data at 1.4 GHz from the NRAO VLA Sky Survey (NVSS, Condon et al. 1996) became available and were used to supplement our spectra. \\subsubsection{WSRT Observations at 1.4 and 5 GHz} The WSRT was used to observe the candidate GPS sources at 1.4 and 5 GHz. The 1.4 GHz observations were performed on 20 February and 10 March 1994 using 8 bands of 5 MHz between 1377.5 and 1423.5 MHz, providing a total bandwidth of 40 MHz. The sources were all observed for about 100 seconds at two to three different hour angles. This resulted in a noise level of typically 1 mJy/beam and a resolution of $15''\\times 15''cosec \\ \\delta$. The results are shown in column 8 of table \\ref{canGPS}. In order to improve the 5 GHz Greenbank flux density measurements, observations were carried out with the WSRT at 4.87 GHz on May 15 1994 using a bandwidth of 80 MHz, at a time when the WSRT was participating a VLBI session. Unfortunately only three telescopes were equipped with 5 GHz receivers. Only sources between 4 and 8 hours right ascension were observed, and the uncertainty in the measured flux densities is large ($\\sim 15$\\%). The resulting flux densities are listed in column 13 in table \\ref{canGPS}. \\begin{table*} \\renewcommand{\\arraystretch}{0.90} \\setlength{\\tabcolsep}{1mm} \\begin{tabular}{|c|rrrrrr|rrr|rr|rrr|c|rrr|} \\hline Source&\\multicolumn{3}{c}{R.A.(1950)}& \\multicolumn{3}{c|}{Decl.(1950)}&$S_{325}$&$S_{609}$&$S^{gb}_{5.0}$&$\\alpha ^{325}_{609}$&$\\alpha ^{325}_{5000}$& $S^{wsrt}_{1.4}$&$S^{VLA}_{8.6}$&$S^{VLA}_{14.9}$&GPS&$S^{nvss}_{1.4}$& $S^{wsrt}_{5.0}$&$S^{merlin}_{5.0}$\\\\ & h & m & s &$^{\\circ}$&$'$&$''$&{\\tiny (mJy)}&{\\tiny(mJy)}& {\\tiny(mJy)}&& &{\\tiny(mJy)}&{\\tiny(mJy)}&{\\tiny(mJy)}& &{\\tiny(mJy)}&{\\tiny(mJy)}&{\\tiny(mJy)}\\\\ \\hline B0400+6042 & 4 & 0 & 7.22 & 60 & 42 & 29.0 & 81 & & 100& &+0.08& 180 & 85& 36 &+& 166& 73& 85 \\\\ B0402+6442 & 4 & 2 & 56.73 & 64 & 42 & 52.4 & 48 & & 69& &+0.13& 63 & 56& 45 & & 70& 32& \\\\ B0406+7413 & 4 & 6 & 37.39 & 74 & 13 & 22.5 & 63 & & 66& &+0.02& 56 & 54& 36 & & 60& 49& \\\\ B0418+6724 & 4 & 18 & 50.86 & 67 & 24 & 7.2 & 47 & & 95& &+0.26& 54 & 96& 79 & & 50& 46& \\\\ B0436+6152 & 4 & 36 & 15.80 & 61 & 52 & 10.0 & 70 & & 127& &+0.22& 208 &108& 60 &+& 238&101&122 \\\\ B0441+5757 & 4 & 41 & 53.90 & 57 & 57 & 21.7 & 60 & & 115& &+0.24& 95 &128& 96 &+& 91& 91&101 \\\\ B0456+7124 & 4 & 56 & 0.15 & 71 & 24 & 10.1 & 88 & & 148& &+0.19& 121 &213&216 & & 116&181& \\\\ B0507+6840 & 5 & 7 & 4.48 & 68 & 40 & 44.7 & 27 & & 33& &+0.07& 30 & 41& 29 & & 31& 28& \\\\ B0513+7129 & 5 & 13 & 38.82 & 71 & 29 & 55.1 &121 & & 181& &+0.15& 236 & 96& 60 &+& 244&123&131 \\\\ B0515+6129 & 5 & 15 & 19.62 & 61 & 28 & 59.3 & 40 & & 45& &+0.04& 26 & 62& 41 & & 28& 43& \\\\ B0518+6004 & 5 & 18 & 42.78 & 60 & 4 & 55.9 & 88 & & 113& &+0.09& 67 & 92&102 & & 101& 56& \\\\ B0531+6121 & 5 & 31 & 55.43 & 61 & 21 & 31.0 & 18 & & 38& &+0.27& 19 & 44& 23 &+& 22& 33& 23 \\\\ B0535+6743 & 5 & 35 & 57.15 & 67 & 43 & 49.8 & 83 & & 182& &+0.29& 97 &235&136 &+&148&177 &168 \\\\ B0536+5822 & 5 & 36 & 8.26 & 58 & 22 & 3.8 & 26 & & 27& &+0.01& 36 & 54& 39 & & 31& 37 & \\\\ B0537+6444 & 5 & 37 & 15.12 & 64 & 45 & 3.7 & 18 & & 32& &+0.21& 29 & 17& 9 &+& 28& 17 & 17 \\\\ B0538+7131 & 5 & 38 & 38.30 & 71 & 31 & 20.9 & 19 & & 77& &+0.51& 45 & 68& 29 &+& 48& 90 & 73 \\\\ B0539+6200 & 5 & 39 & 54.51 & 62 & 0 & 2.2 & 47 & & 104& &+0.29& 123 & 80& 66 &+&126&112 & 99 \\\\ B0542+7358 & 5 & 42 & 50.98?& 73 & 58 & 32.5 & 29 & & 29& &+0.00& 46 & 33& 24 & & 44& 34 & \\\\ B0543+6523 & 5 & 43 & 40.36 & 65 & 23 & 24.6 & 26 & & 43& &+0.18& 65 & 47& 27 &+& 72& 49 & 43 \\\\ B0544+5847 & 5 & 44 & 3.18 & 58 & 46 & 55.8 & 33 & & 42& &+0.09& 67 & 43& 22 &+& 60& 48 & 34 \\\\ B0552+6017 & 5 & 52 & 35.07 & 60 & 17 & 30.1 & 15 & & 26& &+0.28& 44 & 11&$<3$&+& 47& 11 & 13 \\\\ B0556+6622 & 5 & 56 & 13.00 & 66 & 22 & 57.7 & 22 & & 25& &+0.05& 12 & 30& 20 & & 24& 27 & \\\\ B0557+5717 & 5 & 57 & 31.76 & 57 & 17 & 19.7 & 19 & & 29& &+0.16& 63 & 28& 14 &+& 69& 36 & 30 \\\\ B0601+7242 & 6 & 1 & 57.83 & 72 & 42 & 54.6 & 16 & & 26& &+0.18& 14 & 18& 8 & & 14& 37 & \\\\ B0601+5753 & 6 & 1 & 22.05 & 57 & 53 & 31.8 & 19 & & 162& &+0.83& 141 &149&138 &+& &207 &192 \\\\ B0605+7218 & 6 & 5 & 6.15 & 72 & 18 & 51.1 & 24 & & 116& &+0.51& 80 & 39& 60 & & 60& 65 & \\\\ B0607+7335 & 6 & 7 & 33.86 & 73 & 35 & 53.0 & 20 & & 54& &+0.36& 73 & 86& 54 & & 53& 87 & \\\\ B0607+7107 & 6 & 7 & 54.42 & 71 & 8 & 14.1 & 21 & & 25& &+0.10& 24 & 21& 18 & & 27& 42 & \\\\ B0609+7259 & 6 & 9 & 14.93 & 72 & 59 & 50.7 & 20 & & 25& &+0.08& 16 & 16& 11 & & 22& 23 & \\\\ B0738+7043 & 7 & 38 & 37.86 & 70 & 43 & 9.2 & 47 & & 84& &+0.19& 34 & 18&106 & & 73&100 & \\\\ B0741+7213 & 7 & 41 & 30.37 & 72 & 12 & 58.5 & 65 & & 106& &+0.18& 96 & 54& 55 & & 99& 75 & \\\\ B0748+6343 & 7 & 48 & 27.42 & 63 & 43 & 31.7 & 24 & & 54& &+0.42& 39 & 69& 85 &+& 44& 87 &130 \\\\ B0752+6355 & 7 & 52 & 21.41 & 63 & 55 & 59.5 & 15 & & 254& &+1.04& 133 &298&282 &+&196&376 &303 \\\\ B0755+6354 & 7 & 55 & 20.66 & 63 & 54 & 25.4 & 25 & & 38& &+0.15& 26 & 17& 18 &+& 26& 21 & 19 \\\\ B0756+6647 & 7 & 56 & 47.89 & 66 & 47 & 27.8 & 44 & & 164& &+0.48& 90 & 80& 77 &+&104&109 & 97 \\\\ B0758+5929 & 7 & 58 & 13.00 & 59 & 29 & 56.9 & 97 & & 178& &+0.22& 203 &127&101 &+&207&185 &163 \\\\ B0759+6557 & 7 & 59 & 12.95 & 65 & 57 & 45.9 & 15 & & 27& &+0.22& 40 & 14& 8 &+& 48& 26 & 21 \\\\ B0800+6754 & 8 & 0 & 6.92 & 67 & 54 & 31.9 & 78 & & 78& &+0.00& 30 & 33& 32 & & 39& 38 & \\\\ B0802+7323 & 8 & 2 & 32.27 & 73 & 23 & 53.3 &318 & & 321& &+0.00& 277 &387&445 & &307&437 & \\\\ B0808+6518 & 8 & 8 & 5.14 & 65 & 18 & 10.8 & 35 & & 42& &+0.07& 52 & 17& 21 & & 31& 32 & \\\\ B0810+6440 & 8 & 10 & 7.59 & 64 & 40 & 29.1 & 96 & & 193& &+0.26& 92 &132&179 & & 90&133 & \\\\ B0820+7403 & 8 & 20 & 43.56 & 74 & 2 & 53.8 &104 & & 108& &+0.01& 81 & 65& 63 & &102& 97 & \\\\ B0824+6446 & 8 & 24 & 31.62 & 64 & 46 & 27.4 & 24 & & 35& &+0.14& 30 & 11& 13 & & 44& 29 & \\\\ B0826+7045 & 8 & 26 & 52.55 & 70 & 45 & 44.1 & 34 & & 109& &+0.43& 73 & 63& 56 &+& 79& 99 & 92 \\\\ B0827+6231 & 8 & 27 & 3.37 & 62 & 31 & 45.9 & 28 & & 28& &+0.00& 32 & 25& 33 & & 34& 32 & \\\\ B0828+5756 & 8 & 28 & 33.49 & 57 & 56 & 5.6 & 29 & & 32& &+0.04& 37 & 27& 26 & & 53& 40 & \\\\ B0828+7307 & 8 & 28 & 49.08 & 73 & 6 & 58.7 & 81 & & 104& &+0.09& 68 & 58& 68 & &102& 86 & \\\\ B0830+5813 & 8 & 30 & 12.71 & 58 & 13 & 38.8 & 39 & & 65& &+0.29& 65 & 31& 23 &+& 59& 43 & 38 \\\\ B0830+6300 & 8 & 30 & 37.77 & 63 & 0 & 8.1 & 26 & & 36& &+0.12& 54 & 50& 54 & & 62& 65 & \\\\ B0830+6845 & 8 & 30 & 59.69 & 68 & 45 & 33.1 & 53 & & 74& &+0.12& 83 &136&124 & & 86&159 & \\\\ B1525+6801 &15 & 25 & 21.12 & 68 & 1 & 48.9 & 90 & & 103& &+0.05& 153 & 54& 29 &+&161& & 91 \\\\ B1529+6741 &15 & 29 & 17.85 & 67 & 41 & 58.6 & 47 & & 55& &+0.06& 19 & 32& 26 & & 35& & \\\\ B1529+6829 &15 & 29 & 45.15 & 68 & 29 & 9.3 & 51 & & 51& &+0.00& 29 & 21& 23 & & 27& & \\\\ B1536+6202 &15 & 36 & 54.56 & 62 & 2 & 56.4 & 16 & & 30& &+0.23& 15 & 34& 24 & & 22& & \\\\ B1538+5920 &15 & 38 & 27.46 & 59 & 20 & 39.0 & 29 & & 73& &+0.34& 47 & 36& 23 &+& 45& & 45 \\\\ B1539+6156 &15 & 39 & 32.28 & 61 & 56 & 1.9 & 34 & & 40& &+0.06& 15 & 54& 36 & & 33& & \\\\ B1542+6139 &15 & 42 & 5.03 & 61 & 39 & 20.9 & 60 & & 129& &+0.28& 86 &114&119 & & 90& & \\\\ B1542+6631 &15 & 42 & 54.19 & 66 & 31 & 18.2 & 54 & & 86& &+0.17& 44 & 81& 84 & & 51& & \\\\ B1550+5815 &15 & 50 & 55.59 & 58 & 15 & 37.5 & 86 & & 362& &+0.53& 157 &214&212 &+& & &237 \\\\ B1551+6822 &15 & 51 & 53.07 & 68 & 22 & 38.7 & 23 & & 34& &+0.14& 49 & 26& 10 &+& 55& & 27 \\\\\\hline \\end{tabular} \\caption{\\label{canGPS} The sample of candidate GPS sources. Column 1 gives the B1950 source name, column 2 the VLA position, column 3, 4 and 5 the flux densities from WENSS at 325 and 609 MHz and of the Greenbank Survey at 5 GHz. Column 6 and 7 give the 325-609 MHz and the 325-5000 MHz spectral indices, column 8, 9 and 10 the flux densities from the WSRT at 1.4 GHz, and from the VLA at 8.4 and 15 GHz. A cross in column 11 indicates whether the source was selected in the final sample. Column 12, 13 and 14 give the NVSS 1.4 GHz, the WSRT 5 GHz and the MERLIN 5 GHz flux densities.} \\end{table*} \\addtocounter{table}{-1} \\begin{table*} \\renewcommand{\\arraystretch}{0.90} \\setlength{\\tabcolsep}{1mm} \\begin{tabular}{|c|rrrrrr|rrr|rr|rrr|c|rrr|} \\hline Source&\\multicolumn{3}{c}{R.A.(1950)}& \\multicolumn{3}{c|}{Decl.(1950)}&$S_{325}$&$S_{609}$&$S^{gb}_{5.0}$&$\\alpha ^{325}_{609}$&$\\alpha ^{325}_{5000}$& $S^{wsrt}_{1.4}$&$S^{VLA}_{8.6}$&$S^{VLA}_{14.9}$&GPS&$S^{nvss}_{1.4}$& $S^{wsrt}_{5.0}$&$S^{merlin}_{5.0}$\\\\ & h & m & s &$^{\\circ}$&$'$&$''$&{\\tiny (mJy)}&{\\tiny(mJy)}& {\\tiny(mJy)}&& &{\\tiny(mJy)}&{\\tiny(mJy)}&{\\tiny(mJy)}& &{\\tiny(mJy)}&{\\tiny(mJy)}&{\\tiny(mJy)}\\\\ \\hline B1557+6220 &15 & 57 & 8.43 & 62 & 20 & 7.4 & 23 & & 37 & &+0.17& 40 & 12& 4 &+& 42& & 18 \\\\ B1559+5715 &15 & 59 & 5.07 & 57 & 15 & 19.2 & 27 & & 47 & &+0.20& 55 & 36& 39 & & 68& & \\\\ B1600+5714 &16 & 0 & 8.62 & 57 & 14 & 18.8 & 29 & & 45 & &+0.16& 24 & 12& 76 & & 41& & \\\\ B1600+7131 &16 & 0 & 57.00 & 71 & 31 & 40.2 & 26 & & 103 & &+0.50&311 & 37& 15 &+&308& & 85 \\\\ B1604+5939 &16 & 4 & 56.32 & 59 & 39 & 43.7 & 71 & & 110 & &+0.16&111 &130&112 & & & & \\\\ B1607+6026 &16 & 7 & 30.32 & 60 & 26 & 34.6 & 79 & & 79 & &+0.00& 16 & 31& 22 & & 34& & \\\\ B1608+6540 &16 & 8 & 50.55 & 65 & 40 & 15.1 & 66 & & 90 & &+0.11& 63 & 40& 78 & & 68& & \\\\ B1616+6428 &16 & 16 & 26.42 & 64 & 28 & 7.7 & 19 & 37& 62 &+1.06&+0.43& 64 & 58& 58 & & 61& & \\\\ B1620+6406 &16 & 20 & 46.19 & 64 & 6 & 12.6 & 27 & 30& 25 &+0.16&$-$0.03& 41 & 6&$<3$&+& 43& & 11 \\\\ B1622+6630 &16 & 22 & 50.52 & 66 & 30 & 52.6 & 23 & 61& 517 &+1.55&+1.14&178 &230&176 &+&159& &230 \\\\ B1623+6859 &16 & 23 & 36.01 & 68 & 59 & 46.0 & 32 & 32& $<25$&+0.00& & 11 & 3&$<3$& & 18& & \\\\ B1624+6622 &16 & 24 & 7.26 & 66 & 22 & 6.7 & 38 & 42& $<25$&+0.16& & 33 & 4&$<3$& & 26& & \\\\ B1633+6506 &16 & 33 & 7.49 & 65 & 6 & 52.4 & 48 & 66& 114 &+0.51&+0.30&111 & 97& 90 & & 95& & \\\\ B1639+6711 &16 & 39 & 10.76 & 67 & 11 & 47.2 & 34 & 61& 30 &+0.93&$-$0.05& 54 & 27& 19 &+& 76& & 40 \\\\ B1642+6701 &16 & 42 & 16.42 & 67 & 1 & 22.6 &124 & 126& 65 &+0.02&$-$0.24&121 & 43& 24 &+&126& & 56 \\\\ B1645+6738 &16 & 45 & 38.00 & 67 & 38 & 0.8 & 22 & 22& $<25$&+0.00& & 28 & 8& 5 & & 28& & \\\\ B1647+6225 &16 & 47 & 31.26 & 62 & 25 & 49.7 & 31 & 59& 33 &+1.02&+0.02& 69 & 13& 3 &+& 56& & 17 \\\\ B1655+6446 &16 & 55 & 21.09 & 64 & 46 & 21.2 & 23 & 52& 34 &+1.30&+0.14& 68 & 16& 9 &+& 61& & 23 \\\\ B1657+5826 &16 & 57 & 15.96 & 58 & 26 & 31.5 & 56 & 63& 17 &+0.19&$-$0.44& 44 & 18& 13 &+& 48& & 23 \\\\ B1711+6031 &17 & 11 & 39.99 & 60 & 31 & 45.3 & 15 & 29& $<25$&+1.05& & 17 & 3&$<3$& & 19& & \\\\ B1712+6727 &17 & 12 & 50.73 & 67 & 27 & 10.2 & 28 & 30& $<25$&+0.11& & 28 & 23& 13 & & 31& & \\\\ B1714+5819 &17 & 14 & 56.27 & 58 & 19 & 16.2 & 21 & 35& $<25$&+0.81& & 25 & 6& 3 & & 28& & \\\\ B1718+6024 &17 & 18 & 18.75 & 60 & 24 & 11.7 & 19 & 29& $<25$&+0.67& & 20 & 3&$<3$& & 18& & \\\\ B1730+6027 &17 & 30 & 15.71 & 60 & 27 & 24.9 & 41 & 83& 34 &+1.12&$-$0.13&178 &103& 89 & &161& & \\\\ B1746+6921 &17 & 46 & 53.21 & 69 & 21 & 33.5 & 65 & 96& 144 &+0.62&+0.31&161 &127&100 &+&154& &139 \\\\ B1749+6919 &17 & 49 & 31.62 & 69 & 19 & 41.3 & 21 & 31& 18 &+0.62&$-$0.06& 27 & 10& 7 & & 32& & \\\\ B1755+6905 &17 & 55 & 42.48 & 69 & 5 & 48.4 & 16 & 72& 77 &+2.40&+0.28& 84 & 51& 48 & & 78& & \\\\ B1807+5959 &18 & 7 & 17.36 & 59 & 59 & 26.5 & 16 & 43& 30 &+1.52&+0.28& 47 & 37& 22 &+& 42& & 38 \\\\ B1807+6742 &18 & 7 & 23.43 & 67 & 42 & 22.3 & 29 & 52&$<25$ &+0.93& & 47 & 12& 8 &+& 43& & 20 \\\\ B1808+6813 &18 & 8 & 25.41 & 68 & 13 & 36.1 & 33 & 37& 24 &+0.18&+0.04& 42 & 11& 8 &+& 42& & 19 \\\\ B1818+6445 &18 & 18 & 24.79 & 64 & 45 & 17.1 & 35 & 38 &$<25$&+0.13 & & 24 & 27 &$<3$& & 27& & \\\\ B1818+6249 &18 & 18 & 50.02 & 62 & 49 & 56.2 & 41 & 50 &$<25$&+0.32 & & 34 & 8 & 6& & 33& & \\\\ B1819+6707 &18 & 19 & 48.42 & 67 & 7 & 20.8 & 265 &330 & 154 &+0.35 &$-0.20$ & 297 & 93 & 68&+ &311& &142 \\\\ B1821+6251 &18 & 21 & 20.03 & 62 & 51 & 52.6 & 30 & 38 & 31 &+0.38 &+0.01 & 32 & 28 & 19& & 29& & \\\\ B1827+6432 &18 & 27 & 55.40 & 64 & 32 & 13.7 & 152 &228 & 262 &+0.65 &+0.20 & 204 &135 &120& &216& & \\\\ B1829+6419 &18 & 29 & 16.62 & 64 & 19 & 23.1 & 62 & 95 &$<25$&+0.68 & & 80 & 6 & 4& & 75& & \\\\ B1834+6319 &18 & 34 & 48.26 & 63 & 19 & 49.6 & 37 & 46 &$<25$&+0.35 & & 36 & 5 &$<3$& & 36& & \\\\ B1838+6239 &18 & 38 & 12.00 & 62 & 39 & 56.2 & 15 & 33 &$<25$&+1.25 & & 54 & 7 & 6& & 36& & \\\\ B1841+6715 &18 & 41 & 7.21 & 67 & 15 & 51.2 & 36 & 94 & 163 &+1.53 &+0.55 & 142 & 98 &68& + &178& &125 \\\\ B1841+6343 &18 & 41 & 18.25 & 63 & 43 & 56.3 & 15 & 29 &$<25$&+1.05 & & 41 & 10 & 6 & & 36& & \\\\ B1843+6305 &18 & 43 & 6.16 & 63 & 5 & 42.8 & 15 & 41 & 52 &+1.67 &+0.45 & 59 & 27 & 16 & + & 81& & 40 \\\\ B1850+6447 &18 & 59 & 27,78 & 64 & 47 & 31.6 & 49 & 70 &$<25$&+0.57 & & 52 & 6 & $<3$& & 55& & \\\\ B1916+6817 &19 & 16 & 37.66 & 68 & 17 & 51.6 & 23 & 39 &$<25$&+0.84 & & 20 & 6 & 4 & & 26& & \\\\ B1919+6912 &19 & 19 & 57.99 & 69 & 12 & 26.2 & 21 & 28 & 18 &+0.46 &$-0.06$ & 16 & 16 & 14 15& & & & \\\\ B1926+6111 &19 & 26 & 49.66 & 61 & 11 & 20.9 & 404 & & 613 & &+0.15 & 718 & 85 &678 & & &535& \\\\ B1934+7111 &19 & 34 & 41.81 & 71 & 11 & 10.4 & 92 & & 108 & &+0.06 & 142 &119 &102 & &179 & & \\\\ B1938+5824 &19 & 38 & 50.57 & 58 & 24 & 49.9 & 24 & & 29 & &+0.07 & 34 & 30 & 25 & &21 & & \\\\ B1942+7214 &19 & 42 & 2.24 & 72 & 14 & 31.9 & 81 & & 158 & &+0.24 & 233 &147 &110 & + &233 & &183 \\\\ B1944+6007 &19 & 44 & 21.42 & 60 & 7 & 40.5 & 20 & & 79 & &+0.50 & 12 & 62 & 32 & &17 & & \\\\ B1945+6024 &19 & 45 & 24.83 & 60 & 24 & 12.6 & 25 & & 80 & &+0.43 & 55 &125 &188 & + & 55& & 84 \\\\ B1946+7048 &19 & 46 & 12.02 & 70 & 48 & 21.6 & 234 & & 643 & &+0.37 & 887 &389 &268 & + & 953& &574 \\\\ B1951+6915 &19 & 51 & 34.02 & 69 & 15 & 6.3 & 23 & & 32 & &+0.12 & 24 & 40 & 35 & & 33& & \\\\ B1951+6453 &19 & 51 & 42.52 & 64 & 53 & 56.1 & 111 & & 103 & &+0.00 & 89 & 83 & 59 & & 88& & \\\\ B1954+6146 &19 & 54 & 11.69 & 61 & 45 & 58.1 & 66 & & 132 & &+0.25 & 66 &182 &152 & + & 61& &153 \\\\ B1958+6158 &19 & 58 & 45.58 & 61 & 58 & 27.1 & 52 & & 140 & &+0.36 & 111 & 96 & 84 & + & 129& &136 \\\\ B2006+5916 &20 & 06 & 52.14 & 59 & 16 & 43.5 & 30 & & 37 & &+0.08 & 36 & 21 & 18 & & 30& & \\\\ B2011+7156 &20 & 11 & 22.95 & 71 & 56 & 9.4 & 48 & & 134 & &+0.38 & 118 &103 &102 & & & & \\\\\\hline \\end{tabular} \\caption{{\\it Continued...}} \\end{table*} \\subsubsection{VLA Observations at 8.4 and 15 GHz} The candidate GPS sources were observed with the VLA in B-configuration at 8.4 and 15 GHz on 23 July 1994. At both frequencies, the objects were observed in a standard way using a bandwidth of $2\\times 25$ MHz. The phases were calibrated using standard nearby VLA phase calibrators. Total integration times were typically 100 seconds at both frequencies, resulting in noise levels of 0.2 and 1.0 mJy/beam respectively. Systematic errors in flux density of VLA observations at these frequencies are typically about $3\\%$ (eg. Carilli et al. 1991). The data were reduced using AIPS in a standard manner, including several iterations of phase self-calibration. The synthesized beams have half widths of $1.5''\\times 0.8''$ and $0.8''\\times 0.5 ''$ at 8.4 and 15 GHz respectively. Several candidate GPS sources had already been observed at 8.4 GHz on February 26 1994 and April 3 1994 during the Cosmic Lens All Sky Survey (CLASS) program (eg. Myers et al. 1995); these were not re-observed by us at 8.4 GHz. The CLASS 8.4 GHz observations were made using the VLA in A configuration in a standard way, also with a bandwidth of $2\\times 25$ MHz and an average integration time of 30 seconds. The resolution of the CLASS observations was $\\sim 0.2''$, and the noise level $\\sim 0.4$ mJy/beam. The results of the VLA observations are listed in columns 9 and 10 in table \\ref{canGPS}. All of the sources were unresolved, except for B1608+6540, which was found to be a quadruple gravitational lens (Snellen et al. 1995b, Myers et al. 1995, Fassnacht et al. 1996) \\subsubsection{The NRAO VLA Sky Survey at 1.4 GHz} Observations for the 1.4 GHz NRAO VLA Sky Survey (NVSS, Condon et al. 1996) began in September 1993 and are planned to cover the sky north of declination $-40^{\\circ}$ (82\\% of the celestial sphere). Data in our regions of interest were taken on 1 November 1993 for the region $4^h00^m15$& $>188$& 0.54& 0.70& -\\\\ B1946+7048& 1.8& 929& 0.91& -0.64& 0.6\\\\ B1954+6146& 8.4& 169& 0.00& -0.31& 1.4\\\\ B1958+6158& 3.3& 142& 0.52& -0.23& 0.9\\\\ \\hline \\end{tabular} \\end{center} \\caption{\\label{GPS}The resulting sample of GPS sources.} \\end{table} \\begin{figure*} \\vspace{-0.5cm} \\hbox{\\hspace{-0.8cm} \\psfig{figure=snellen0400+6042.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen0436+6152.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen0441+5757.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen0513+7129.ps,width=5.1cm} } \\vspace{-0.5cm} \\hbox{\\hspace{-0.8cm} \\psfig{figure=snellen0531+6121.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen0535+6743.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen0537+6444.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen0538+7131.ps,width=5.1cm} } \\vspace{-0.5cm} \\hbox{\\hspace{-0.8cm} \\psfig{figure=snellen0539+6200.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen0543+6523.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen0544+5847.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen0552+6017.ps,width=5.1cm} } \\vspace{-0.5cm} \\hbox{\\hspace{-0.8cm} \\psfig{figure=snellen0557+5717.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen0601+5953.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen0748+6343.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen0752+6355.ps,width=5.1cm} } \\vspace{-0.5cm} \\hbox{\\hspace{-0.8cm} \\psfig{figure=snellen0755+6354.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen0756+6647.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen0758+5929.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen0759+6557.ps,width=5.1cm} } \\caption{\\label{spectra} Radio spectra of individual sources. Crosses indicate MERLIN data, diamonds indicate at 1.4 GHz NVSS data and at 5 GHz WSRT data.} \\end{figure*} \\addtocounter{figure}{-1} \\begin{figure*} \\vspace{-0.5cm} \\hbox{\\hspace{-0.8cm} \\psfig{figure=snellen0826+7045.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen0830+5813.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1525+6801.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1538+5920.ps,width=5.1cm} } \\vspace{-0.5cm} \\hbox{\\hspace{-0.8cm} \\psfig{figure=snellen1550+5815.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1551+6822.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1557+6220.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1600+7131.ps,width=5.1cm} } \\vspace{-0.5cm} \\hbox{\\hspace{-0.8cm} \\psfig{figure=snellen1620+6406.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1622+6630.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1639+6711.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1642+6701.ps,width=5.1cm} } \\vspace{-0.5cm} \\hbox{\\hspace{-0.8cm} \\psfig{figure=snellen1647+6225.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1655+6446.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1657+5826.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1746+6921.ps,width=5.1cm} } \\vspace{-0.5cm} \\hbox{\\hspace{-0.8cm} \\psfig{figure=snellen1807+5959.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1807+6742.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1808+6813.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1819+6707.ps,width=5.1cm} } \\caption{{\\it Continued...}} \\end{figure*} \\addtocounter{figure}{-1} \\begin{figure*} \\vspace{-0.5cm} \\hbox{\\hspace{-0.8cm} \\psfig{figure=snellen1841+6715.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1843+6305.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1942+7214.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1945+6024.ps,width=5.1cm} } \\vspace{-0.5cm} \\hbox{\\hspace{-0.8cm} \\psfig{figure=snellen1946+7048.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1954+6146.ps,width=5.1cm}\\hspace{-0.8cm} \\psfig{figure=snellen1958+6158.ps,width=5.1cm} } \\caption{{\\it Continued...}} \\end{figure*} ", "conclusions": "A sample of GPS sources has been selected from the Westerbork Northern Sky Survey, with flux densities one to two orders of magnitude lower than bright GPS sources investigated in earlier studies. Sources with inverted spectra at frequencies $>325$ MHz have been observed with the WSRT at 1.4 and 5 GHz and with the VLA at 8.6 and 15 GHz to select genuine GPS sources. This has resulted in a sample of 47 GPS sources with peak frequencies ranging from $\\sim$500 MHz to $>$15 GHz, and peak flux densities ranging from $\\sim$40 to $\\sim$900 mJy. Five GPS sources in our sample show extended emission or nearby components in the NVSS maps at 1.4 GHz. However it is not clear if these components are related to the GPS sources. About 30\\% of the objects show flux density differences greater than 20\\% between the Greenbank and MERLIN 5 GHz measurements, with the Greenbank data points all higher than the MERLIN observations. We believe this is due to variability, and that the lack of sources with reverse variability (the MERLIN flux density greater than the Greenbank flux density) is due to a selection effect caused by the ``old'' epoch (1987) of the Greenbank observations. GPS source counts are comparable to 1/250 of the 2 GHz source counts for large scale radio sources, if the latter sources were to have 10 times their measured flux densities. Unfortunately, apparent differences in redshift distributions between GPS and large scale radio sources hamper a direct and straightforward interpretation of the source counts. Potentially, the comparison of GPS source counts with that of large scale radio sources can provide clues about the age of GPS sources and their luminosity evolution. If it is assumed that the redshift distributions are the same for GPS and large size radio sources, the source counts indicate that GPS sources have to decrease in luminosity by a factor of $\\sim 10$ if they all evolve into large scale radio sources." }, "9803/astro-ph9803283_arXiv.txt": { "abstract": "We present the results of a numerical code that combines multi-zone chemical evolution with 1-D hydrodynamics to follow in detail the evolution and radial behaviour of gas and stars during the formation of elliptical galaxies. We use the model to explore the links between the \\ev\\ and formation of elliptical galaxies and QSO activity. The knowledge of the radial gas flows in the galaxy allows us to trace metallicity gradients, and, in particular, the formation of a high-metallicity core in ellipticals. The high-metallicity core is formed soon enough to explain the metal abundances inferred in high-redshift quasars. The star formation rate and the subsequent feedback regulate the episodes of wind, outflow, and cooling flow, thus affecting the recycling of the gas and the chemical enrichment of the intergalactic medium. The \\ev\\ of the galaxy shows several stages, some of which are characterized by a complex flow pattern, with inflow in some regions and outflow in other regions. All models, however, exhibit during their late \\ev\\ a \\gw\\ at the outer boundary and, during their early \\ev, an inflow towards the \\gal\\ nucleus. The characteristics of the inner inflow could explain the bolometric luminosity of a quasar lodged at the galaxy centre as well as the evolution of the optical luminosity of quasars. ", "introduction": "Imaging studies of the faint extensions around QSOs indicate that \\el\\ \\gals\\ are the host \\gals\\ of the radio-loud and the brightest QSOs (Smith et al. 1986; Hutchings, Janson \\& Neff 1989; Hutchings et al. 1994; Hutchings \\& Morris 1995; McLeod \\& Rieke 1995; Aretxaga, Boyle \\& Terlevich 1995; Disney et al. 1995; Bahcall, Kirkakos \\& Schneider 1996; Ronnback et al. 1996; Taylor et al. 1996). As a matter of fact, at high redshifts ($z > 2$), the only galactic systems available to harbour QSOs are the spheroids, since the disks are formed much later. In addition, the epoch of completion of the large spheroids ($z > 2$) coincides with the peak in the QSO activity ($2 < z < 3$) (Schmidt et al. 1991), suggesting a relation between the QSO phenomenon and the \\for\\ of large \\el s. In fact, in recent years there has been increasing evidence linking QSO activity with galaxy formation. The high metal content of high redshift QSOs, the high dust content (several $10^8$ \\msun\\ of dust) of distant QSOs (Andreani, Franca \\& Cristiani 1993; Isaak et al. 1994; McMahon et al. 1994; Omont et al. 1996) plus the possible relation between the galaxy luminosity function (LF) and the QSO LF (Terlevich \\& Boyle 1993, hereafter TB93; Boyle \\& Terlevich 1998, hereafter BT98) provides tantalising evidence of this link. In any case, the fact that even the highest redshift QSO has strong metal lines in its spectrum requires that the broad line region (BLR) gas has been enriched by a stellar population formed before $z\\sim 5$. Work by TB93 and Hamann \\& Ferland (1993) (hereafter HF93) highlighted the importance of metal production in the early evolution of a galaxy. Since QSOs are seen up to redshifts of $z\\sim 5$, the supersolar metallicities required by the BLR models should be reached by $\\sim 1$~Gyr since the beginning of the galaxy formation epoch. This evolutionary time scale is an important constraint in chemical enrichment models. Two fundamental relations involving intrinsic parameters of elliptical galaxies show remarkable little dispersion and point towards an early formation of ellipticals. They are the colour-luminosity relation and the ``fundamental plane'' relating the total luminosity of an elliptical to its central velocity dispersion and surface brightness. The tightness of the colour-luminosity relation provides strong evidence that most of the present stellar population was formed at $z > 2$ (Bower, Lucey and Ellis 1992). The narrowness of the ``fundamental plane'' gives additional support to that conclusion (Renzini and Ciotti 1993) and in addition indicates that the properties of the core (velocity dispersion, Mg$_2$ strength) are intimately linked to the galaxy global ones (D$_n$, luminosity). There is also good evidence that the stellar population in massive ellipticals is metal rich with respect to the Sun, and shows large radial metallicity gradients (e.g., Worthey, Faber and Gonzalez, 1992; Davies, Sadler and Peletier 1993). These properties of the \\el\\ \\gals\\ may also be related to the fact that they probably harbour QSOs in their centres at some stage of their \\ev. One-zone chemical evolution models have been used (Hamann \\& Ferland 1992, HF93, Padovani \\& Matteucci 1993, Matteucci \\& Padovani 1993) to investigate the chemical history and the fueling of QSOs. However, during the early evolution of the elliptical galaxy, it is expected several episodes of gas outflow and inflow which cannot be followed by the one-zone model. In addition, the one-zone chemical evolution models that attempt to explain the high metal content in high redshift QSOs tend to overproduce metals (averaged over the entire galaxy) and predict an excessively high luminosity for the parent galaxy of the QSO. For example, HF93 model M4 reproduces the rapid metal production needed but it is overluminous. In this model, an elliptical of $10^{11}$ \\msun\\ has a peak bolometric luminosity of $\\sim 2 \\times10^{13}$ \\lsun\\ at an age of $\\sim 0.1$ Gyr. But an elliptical with $10^{11}$ \\msun\\ is at present only a sub-$L^*$ galaxy (with a blue luminosity of $\\sim 0.3 L^*$, for $M_B^*=-21$ and $[M/L_B]=10$), so that for the most luminous ($M_B \\approx -24$) ellipticals in the nearby Universe HF93 model M4 predicts luminosities of up to $10^{15}$ \\lsun\\ during its formation. These luminosities are higher than the QSO luminosities!! Note that only the core of the galaxy has to be metal rich in accordance with the observed metallicity gradients in nearby galaxies. In the starburst model for QSOs (TB93 and references therein), the QSOs are the young cores of massive ellipticals forming most of the dominant metal-rich population in a short starburst. The core mass, which participates in the starburst, comprises only a small fraction ($\\sim 5$ \\%) of the total galaxy mass. Also in the standard supermassive black hole model for QSOs, only $0.5-1$ per cent of the \\gal\\ mass goes into the black hole (Haehnelt \\& Rees 1993). The excessive production of energy and metals in the one-zone model arises, therefore, from its inability to resolve the core of the \\gal. The QSO LF undergoes strong \\ev\\ between $z=2$ and the present epoch, with the redshift dependence of the LF being well-described by a constant comoving space density and a pure power-law luminosity \\ev\\ $L(z) \\propto (1+z)^k$ (Boyle et al. 1988, 1991, BT98), with $k$ in the range $3.11$ Gyr is consistent with the decline of luminosity inferred for $z<2$ from the \\ev\\ of QSO LF with redshift. One of our models exhibits recurrent late central \\cf\\ episodes, which are brief (a few $10^7$ yr) and involve decreasing amounts of mass. The gas inflow into the inner 100 pc is regulated by episodes of star formation leading naturally to several short episodes of central inflow, thus giving support to the episodic scenario for evolution of the LF of QSOs. The model also explains the luminosities of QSOs. The central \\cf\\ rates explain bolometric luminosities of up to $10^{47}$ erg s$^{-1}$, for an efficiency $f$ of mass-energy conversion of 0.1. However, the bolometric luminosities of the brightest QSOs ($\\approx 10^{48}$ erg s$^{-1}$) cannot be explained by a continuous deposition of the central inflow. Rather, the highest QSO luminosities require a discrete deposition, in which the gas is accumulated during $\\approx 1$ Gyr and then consumed by a central engine in few $10^7$ yr. Accordingly, we have made some predictions on the QSO LF at $z \\ga 1$ based on a simple discontinuous model for QSO activity, in which there is two short gas consumption events during the first central inflow episode. We scaled the QSO LF to the present day elliptical LF, assuming that all \\el s have harboured a QSO during their \\ev. Both the starburst and the supermassive black hole models predict the right shape of the QSO LF, but the nuclear starburst systematically underestimates the density number of QSOs. In addition, our model reproduces the \\ev\\ of the LF between $z=1.25$ and $z=2.9$. In our models, the mass deposited by the first central inflow represents 0.15-0.5 \\% of the present day luminous mass of the \\gal. Assuming that this mass goes to the formation of a central object (star cluster or black hole), the model correctly predicts for the Dark Massive Objects (DMOs) in the nuclei of \\gals, both the masses and the DMO-to-galaxy mass ratios. Another conclusion derived from our models is that the hosts of high-redshift AGN should be relatively mature objects. The calculated \\ev\\ of the inner \\cf\\ and energetic considerations imply that the gas deposited by the inflow in the nucleus should accumulate during $\\sim 1$ Gyr before triggering a short-lived AGN activity event. On the other hand, except for extremely large \\gals\\ (present-day $M_B=-24$), the first, massive \\cf, responsible for maintaining the AGN activity, lasts for $2-3$ Gyr. In this scenario, if the high redshift \\gal\\ is to display strong AGN activity, its probable age would range from $\\sim 1$ to $\\sim 3$ Gyrs. The minimum age of $\\sim 1$ Gyr for QSO hosts derived above is consistent with the $\\sim 1$ Gyr time scale for metal enrichment, needed to explain the strong metal lines observed even in $z\\sim 5$ QSOs. One further prediction of our models is that, at the present, only the most massive objects should be host to powerful AGN, but at high redshift powerful AGN activity is expected even for hosts of smaller mass. Interestingly enough, this is what seems to be observed, for radio \\gals\\ at least. Models 1 and 2 are examples of sub-$L^*$ \\gals\\ with strong AGN activity at high redshift. Note that $\\zeta$ increases with decreasing \\gal\\ mass for masses below $M_G \\approx 10^{12}$ \\msun. Since this parameter describes the relative importance of the first \\cf\\ episode occuring when the \\gal\\ is less than $2-3$ old, this means that for the lower mass systems, the efficiency of building-up the nucleus is higher than in larger systems, and that they have an early \\cf\\ massive enough to sustain a strong AGN activity. However, this activity is limited only to the 2-3 first Gyr of the \\gal\\ and, therefore, smaller systems with strong nuclear activity are to be found only at high redshift. Even in the case of the episodic scenario of model 2(1/3), the late nuclear activity is much weaker at low redshift (in this model, the two late central inflow episodes deposit only 1.8 \\% and 1 \\% of the mass of the first inflow episode). On the other hand, models 10, 20 and 50 exhibit a present-day massive central \\cf\\ which could trigger AGN activity. In these massive systems, the late cooling flow may accumulate into the nucleus an amount of mass comparable to that of the early \\cf\\ (model 20 is typical of this case, the late \\cf\\ deposits $1.65\\times 10^9$ \\msun, i.e. 47 \\% of the mass of the first \\cf). Only these objects, therefore, would harbour, intense AGN activity at low redshift. Support to this picture is given by recent analysis of host \\gals\\ of powerful nearby ($z \\la 0.3$) AGN, belonging to three samples --- radio \\gals\\, radio-loud quasars, and radio-quiet quasars (Taylor et al. 1996). For all three classes of AGN, the host \\gals\\ are large (half-light radius $r_{1/2} \\geq 10$ kpc) and luminous (K-band luminosity $L_K\\geq L^*$). (Note that the less massive model exhibiting a present-day central is model 10, with $M_B=-21.8$, or $L_B=2 L^*$.) Finally, we should note that, although pure luminosity \\ev\\ seems to reproduce the \\ev\\ of the QSO LF, in our models the individual QSOs do not dim over cosmological time scales, but rather are short-lived (i.e., $t_{on}=1-3\\times 10^7$ yr). In order to comply with the energetic requirements of the QSOs, their activity must occur in short episodes of massive consumption of mass accumulated in the galactic nucleus during a much longer span of time ($\\sim 1$ Gyr). This sort of episodic activity displayed by our model for QSOs seems to be a rule among the AGN in general, since for other class of AGN, the radio galaxies, the radio LF also seems to follow pure luminosity \\ev, and yet the radio sources themselves seem to have lifetimes of only a few $10^7$ yr. As a matter of fact, a comparison between the properties of the radio-loud population and the present model would be very useful to clarify the relation of the QSO phenomenon and the early \\ev\\ of \\el\\ \\gals, since radio observations allow us to explore the $z>2$ domain without the need of uncertain corrections for dust absorption and lensing bias that hamper the optical tecniques." }, "9803/astro-ph9803297_arXiv.txt": { "abstract": "We discuss the chemical evolution of dwarf irregular and blue compact galaxies in light of recent data, new stellar yields and chemical evolution models. We examine the abundance data for evidence of \\hii\\ region self-enrichment effects, which would lead to correlations in the scatter of helium, nitrogen, and oxygen abundances around their mean trends. The observed helium abundance trends show no such correlations, though the nitrogen--oxygen trend does show strong evidence for real scatter beyond observational error. We construct simple models for the chemical evolution of these galaxies, using the most recent yields of \\he4, C, N and O in intermediate- and high-mass stars. The effects of galactic outflows, which can arise both {}from bulk heating and evaporation of the ISM, and from the partial escape of enriched supernova ejecta are included. In agreement with other studies, we find that supernova-enriched outflows can roughly reproduce the observed He, C, N, and O trends; however, in models that fit N versus O, the slopes $\\Delta Y/\\Delta$O and $\\Delta Y/\\Delta$N consistently fall more than $2\\sigma$ below the fit to observations. We discuss the role of the models and their uncertainties in the extrapolation of primordial helium from the data. We also explore the model dependence arising nucleosynthesis uncertainties associated with nitrogen yields in intermediate mass stars, the fate of $8-11 \\msol$ stars, and massive star winds. ", "introduction": "Helium-4 plays a central role in big bang nucleosynthesis. The calculation of the primordial helium abundance is robust, being only weakly sensitive to the cosmic baryon-to-photon ratio, $\\eta$. However, the helium abundance is quite sensitive to, and provides a strong constraint on the physics of the early universe. The primordial abundance of helium (with mass fraction \\yp) is best determined via observations of \\hii\\ regions in the most metal-poor galaxies---dwarf irregulars and blue compact galaxies (hereafter, BCGs). For these systems, the helium evolution is derived empirically: the data show that the helium abundance increases with metallicity indicators such as oxygen and nitrogen. Following Peimbert \\& Torres-Peimbert, \\pcite{ptp}, the primordial abundance of helium is inferred from an extrapolation of the observed trend to zero metallicity. As emphasized by Fields \\& Olive\\pcite{fo}; Fields, Kainulainen, Olive, \\& Thomas \\pcite{fkot}; and Olive, Steigman, \\& Skillman (\\cite{oss}; hereafter OSS97), the empirical extrapolation of primordial helium sidesteps any reliance on detailed results of chemical evolution models, especially since the extrapolation to zero metallicity from the data is small. Indeed, until the measurements of D/H in quasar absorption systems can be confirmed to represent a uniform primordial value, the \\he4 abundance is crucial when used in conjunction with \\li7 as a test of big bang nucleosynthesis theory. (This remains true so long as $\\yp \\la 0.24$; at high \\yp, the insensitivity of \\yp\\ to $\\eta$ makes \\he4 a very poor discriminator for primordial nucleosynthesis.) Thus, given the importance of \\he4 as observed in BCGs, it is clearly of interest to compare the predictions of chemical evolution with the observations of these systems. A less than successful comparison of theory with observation could point to the subtleties in chemical evolution modeling of even these (apparently) simple systems. Several groups have modeled the chemical evolution of BCGs (Matteucci \\& Chiosi \\cite{mc}; Matteucci \\& Tosi \\cite{mt}; Gilmore \\& Wyse \\cite{gw}; Pilyugin \\cite{pil}; Clayton \\& Tantelaki \\cite{cp}; Marconi, Matteucci \\& Tosi \\cite{mmt}; Carigi, Col\\'{\\i}n, Peimbert, \\& Sarimento \\cite{ccps}). The high star formation activity in BCGs indicates that star formation must occur in stochastic bursts, unless the systems are very young; all models included this behavior. In addition, these galaxies probably drive outflows of material, perhaps with efficiencies high enough to significantly alter the chemical and/or dynamical evolution. For example, the gas fraction--metallicity relations of BCGs may not be compatible with closed box models, suggesting the importance of gas outflow via supernova-driven outflows (Lequeux, Rayo, Serrano, Peimbert, \\& Torres-Peimbert \\cite{leq}; Matteucci \\& Chiosi \\cite{mc}; Carigi, Col\\'{\\i}n, Peimbert, \\& Sarimento \\cite{ccps})---though the difficulties of determining an accurate gas--to--baryon fraction lead to significant uncertainties in these arguments. In any case, X-ray observations of diffuse, hot gas clearly support the existence of outflows in some BCGs (Heckman et al.\\ \\cite{heck}; Della Ceca, Griffiths, Heckman, \\& Mackenty \\cite{dc_etal}; Della Ceca, Griffiths, \\& Heckman \\cite{dcgh}), though Skillman \\pcite{evan96} and Skillman \\& Bender \\pcite{sb} present arguments against a dominant role of outflows in {\\em all} dwarf galaxies. Independent of the question of outflows, another key result from these studies showed that the usual instantaneous {\\em mixing} approximation is inappropriate. Self-enrichment of \\hii\\ regions in a burst phase (Kunth \\& Sargent \\cite{ksgt}; Kunth, Lequeux, Sargent, \\& Viallefond \\cite{klsv}; but see Pettini \\& Lipman \\cite{pl}) can lead to significant scatter in abundance trends. Most recently, it has been argued that the due to the intense bursts of star formation and energy release in the evolution of these systems, their stability requires of the presence of significant amounts of dark matter (Bradamante, Matteucci, \\& D'Ercole \\cite{bmd}). This may have an impact on the effect of supernovae driven winds on the elements abundances. In a somewhat different approach, Mac Low \\& Ferrara \\pcite{mlf} take the presence of dark matter as a starting point, and examine the requirements for BCG outflows as a function of supernova rate and galaxy mass. Their models of the the dark matter potential, and the detailed gas dynamics of outflows, suggest that the loss of the entire ISM---``blowaway''---is only possible for smaller galaxies ($\\la 10^6 \\msol$). However, Mac Low \\& Ferrara also find that the preferential escape of the energetic and metal-enriched supernova ejecta is much easier, occurring at significant levels for galaxies with baryonic masses up to $\\sim 10^8 \\msol$. While there has been great progress in laying out the basic features of BCG chemical evolution, some important issues remain unresolved, including some bearing on derivation of primordial helium. For example, there remains some question as to the theoretical justification for the linear fit of helium versus metals, particularly nitrogen (e.g., Mathews, Boyd, \\& Fuller \\cite{mbf}; Balbes, Boyd, \\& Mathews \\cite{bbm}). Closely related to this is the dependence of the inferred \\yp ~on the metallicity tracer. Several studies (e.g., OSS97; Olive, Steigman, \\& Walker \\cite{osw}) have found that the primordial helium abundance derived from a linear fit to $Y-$N consistently differs by about $1\\sigma$ {}from that derived via $Y-$O. While this difference is as yet too mild to be a problem, it does suggest need to re-examine the appropriateness of the linear fits in the context of detailed chemical evolution models. Recent new observations such as the growing set of carbon observations are ripe for more theoretical attention. In this paper, we examine the chemical evolution of BCGs, with particular attention to these issues related to primordial helium, and the uncertainties in the chemical evolution modeling. In particular, we note the interplay between galactic outflow prescriptions and nucleosynthesis uncertainties. Different parameterizations for the nucleosynthesis of nitrogen and their role \\yp(N) versus \\yp(O) discrepancy are also considered. We note the model-dependence of $\\Delta Y/\\Delta Z$ and other slopes, and examine in detail a suggestion by Fields \\pcite{fields} that helium production processes due to low-mass supernovae or high-mass stellar winds may account for the large observed values of the slopes. We conclude that the empirical fitting procedures used to obtain \\yp\\ do find support from our models. However, we also note that models using the most recent and detailed nucleosynthesis yields do no reproduce the observed $Y-$N,O trends, even in the presence of outflows. Possible solutions to this problem are discussed. ", "conclusions": "We have considered the chemical evolution of BCGs, through an analysis of the observed HeNO abundances, and by studying chemical evolution models. Our basic conclusions are as follows. The scatter in the $Y-$N and $Y-$O data is entirely consistent with the observational errors. In particular, these data do not show the correlations (in the N--O plane) that one would expect if the scatter were due to self-enrichment by massive star ejecta. On the other hand, it appears that the N--O scatter real. We expect that at least some of the observed scatter is due to self-enrichment, as is also suggested by the observed correlation of high C/O systems with high N/O. Our study of chemical evolution models of BCGs concludes that the favored models for these systems (i.e., including outflows), coupled with recent and detailed stellar yield calculations (which are metallicity dependent), are not successful in reproducing simultaneously all of the abundance trends. We do find that in general, the models do predict linear slopes in He versus N and O, as has long been assumed in phenomenological fits to the data. The large slopes are, however, hard to reproduce quantitatively---even in models with supernova-enriched outflows. The only models which {\\em can} produce large slopes, while (possibly) maintaining agreement with N--O data, are those using the (less detailed) stellar yields of Maeder \\pcite{mae} in combination with a log-normal IMF. Consequently, we suspect that the root of the problem lies in the nucleosynthesis inputs, particularly the He yields. We emphasize that any means of improving the helium slopes must be tested in models which fit {\\em all} of HeNO, and preferably C as well. Another shortcoming of the models regards nitrogen. We find that the models with the detailed yields typically do not predict a strong enough secondary character for N versus O. As a result, they do not find that the $\\Delta Y/\\Delta $N slopes change much with metallicity. In particular, the models do not reproduce the differences found in the $Y-$N slopes (and intercepts) between the full and low-metallicity data sets. Despite these unresolved issues, our study of BCG models does allow us to draw some conclusions. Most models with detailed yields predicted $Y-$N,O relations that were close to linear; this gives some theoretical justification for the adoption linear regressions of the data (particularly for $Y$--O). Also, we confirm that the nature of the star formation rate---bursting versus smooth---does strongly affect the detailed evolutionary history, and an ensemble of different histories leads to significant scatter in N--O. Fortunately, however, the N--O evolution in the smooth star forming history forms an upper envelope to the N--O trends in the bursting models. Thus, if one is not interested in the scatter, one can adopt a smooth star forming model as long as one is careful to fit appropriate envelope of the data. We have also shown that carbon data provides an important additional constraint on BCG chemical evolution. In particular, C and N together can diagnose the tradeoff between C and N controlled by hot bottom burning. Additional C observations would be very useful in further constraining BCG models. Finally, aspects of our results have implications for systems other than BGCs. The difficulty for our models to reproduce the helium slopes probably carries over to our own Galaxy. Although there is some evidence that the Galactic $\\Delta Y/\\Delta Z$ may be lower than in BCGs, it nevertheless seems to exceed the values ($\\sim 1$) consistently predicted by all of our models without strong enriched winds. A good solar neighborhood $\\Delta Y/\\Delta Z$ would help illuminate whether the helium slope problem is due to yields, or due to environmental differences between the Galaxy and BCGs. Finally, we note the similarity between the low metallicity N--O evolution in BCGs, QSO absorbers (c.f. Lu, Sargent, \\& Barlow \\cite{lsb}), and other external galaxies (van Zee, Salzer, \\& Haynes \\cite{vzsh}). The observed trends are compatible with each other, at least to a first approximation. This suggests that the N--O evolution of these objects is similar, which may then imply that the enriched outflow of BCGs is small. If so, this is additional evidence that the heart of the problem of the low helium slopes lies in the stellar yields themselves." }, "9803/astro-ph9803318_arXiv.txt": { "abstract": "We use high-resolution hydrodynamic simulations to investigate the density profile of hot gas in clusters of galaxies, adopting a variant of cold dark matter cosmologies and employing a cosmological N-body/smoothed particle hydrodynamics code to follow the evolution of dark matter and gas. In addition to gravitational interactions, gas pressure, and shock heating, we include bremsstrahlung cooling in the computation. Dynamical time, two-body relaxation time, and cooling time in the simulations are examined to demonstrate that the results are free from artificial relaxation effects and that the time step is short enough to accurately follow the evolution of the system. In the simulation with nominal resolution of $66h^{-1}\\,{\\rm kpc}$ the computed cluster appears normal, but in a higher (by a factor 2) resolution run, cooling is so efficient that the final gas density profile shows a steep rise toward the cluster center that is not observed in real clusters. Also, the X-ray luminosity of $7\\times10^{45}{\\rm ergs\\,s^{-1}}$ far exceeds that for any cluster of the computed temperature. The most reasonable explanation for this discrepancy is that there are some physical processes still missing in the simulations that actually mitigate the cooling effect and play a crucial role in the thermal and dynamical evolution of the gas near the center. Among the promising candidate processes are heat conduction and heat input from supernovae. We discuss the extent to which these processes can alter the evolution of gas. ", "introduction": "Hot X-ray--emitting gas in clusters of galaxies contains a variety of information relevant to many fields in astrophysics. Density and temperature profiles of the gas give among the most reliable estimates of cluster mass, which are of unparalleled importance to cosmology (e.g., \\cite{bah95}). They will also reflect physical processes that have played a crucial role in the thermal and dynamical evolution of the gas. In addition to gravitation, hydrodynamics, and shock heating, such processes may include radiative cooling, heat conduction, and feedback from star formation. One of the central questions about the cluster gas structure involves the presence of a core. Observations have revealed that the gas density profile has a distinct core, inside of which density is nearly constant. Gravitational N-body simulations have demonstrated that gravity alone cannot produce such a core in an object formed as a result of hierarchical structure formation; halos formed in high-resolution N-body simulations have density profiles with significant slope toward the center up to the resolution limit (\\cite{nav96}, 1997; \\cite{fuk97}; \\cite{moo98}). For example, Navarro, Frenk, and White (1996, 1997) found that the density profiles of halos with a wide range of masses can be fitted by \\begin{equation} \\label{eqn:nfw} \\rho(r) = \\frac{\\rho_c}{(r/r_s)(1+r/r_s)^2} \\end{equation} (hereafter referred to as the NFW profile), with $\\rho_c$ and $r_s$ being the fitting parameters. In the NFW profile, gas temperature would approach zero in a cluster's central regions ($T \\propto r$) were it to have the same profile as dark matter. While this model does produce a convergent X-ray luminosity, the temperature structure is different from what is observed. It follows that processes other than gravity are responsible for the formation of the cores. Makino, Sasaki and Suto (1998) pointed out that the gas distribution develops a core in a dark matter halo having a central cusp (e.g., the NFW profile) if the gas is isothermal and in hydrostatic equilibrium. Even if the gas obeys a less stringent limit of constant entropy ($T/n^{2/3} \\to {\\rm const}$), it would avoid a central cusp and have an apparently constant-density, constant-temperature core. But it is not evident what physical mechanisms could enforce either isothermality or isentropy. Many authors have studied the evolution of clusters using simulations that include hydrodynamics (\\cite{evr90}; \\cite{tho92}; \\cite{kat93}; \\cite{bry94a}, 1994b; \\cite{kan94}; \\cite{nav95}; \\cite{bar96}; \\cite{bry97}, 1998; \\cite{eke98}; \\cite{pen98}; \\cite{yos98}). In general these simulations have succeeded in producing clusters that have cores similar to those observed. However, the situation is far from satisfactory for at least two reasons. First, artificial two-body relaxation may affect the dynamics, especially in the central region. Indeed, Steinmetz and White (1997) showed that this effect gives rise to artificial energy transfer from dark matter to gas. Both spatial and mass resolutions are only marginally adequate in most published works. Second, almost none of the simulations include radiative cooling of gas. Although cooling is probably unimportant in the outer part of a cluster (\\cite{sar86}), it may affect the dynamics in the central region. This work is an investigation of cluster gas using hydrodynamic simulations that attempt to address these limitations. First, we minimize two-body relaxation effects for a given computational cost by employing a multiresolution technique that enables us to improve resolution only inside the clusters where we really need high resolution. Second, we include cooling due to bremsstrahlung, which dominates cooling of gas at above $\\gtrsim 10^7{\\rm K}$. We ignore line cooling, but this is not a serious problem because we focus our attention on X-ray--emitting gas. A practical reason for ignoring line cooling is that doing so enables us to avoid very short timescales in moderate temperature ($10^4$ -- $10^6 {\\rm K}$), high density regions. Allowing for line cooling would strengthen the conclusions of this paper. We organize the rest of the paper as follows. In \\S\\ 2 we describe the method and parameters of the simulations. We present the results in \\S\\ 3. As we will see, the most important result is that cooling and increased resolution give rise to a density profile of the gas that rises steeply toward the center and consequently produces excessive X-ray luminosity. In \\S\\ 4 we discuss the implications of our results in connection with physical processes that are still missing in the simulations. In \\S\\ 5 we give our conclusions. ", "conclusions": "\\label{sect:conclusion} Finally let us briefly summarize our conclusions. \\begin{enumerate} \\item Our multiresolution simulation has followed with reasonable accuracy the evolution of gas and dark matter in a typical cluster under the influence of bremsstrahlung cooling as well as gravity and hydrodynamics. \\item The high resolution simulation resulted in a gas density profile steeply rising toward the center, with consequent very high X-ray luminosity; however, these properties are not observed. \\item Heat conduction and SN heating are among the processes that may account for the discrepancy. Had we allowed for their likely importance in the real world we might have been able to recover the observed gas density profile. In a future work we will examine their effects by directly incorporating them into the simulation. \\end{enumerate}" }, "9803/astro-ph9803154_arXiv.txt": { "abstract": "Recent calculations indicate that in the outer parts of neutron stars nuclei are rod-like or slab-like, rather than roughly spherical. We consider the elastic properties of these phases, and argue that they behave as liquid crystals, rather than rigid solids. We estimate elastic constants and discuss implications of our results for neutron star behavior. ", "introduction": " ", "conclusions": "" }, "9803/astro-ph9803224_arXiv.txt": { "abstract": "We examine all possible stationary, optically thick, geometrically thin accretion disc models relevant for active galactic nuclei (AGN) and identify the physical regimes in which they are stable against the thermal--viscous hydrogen ionization instability. Self--gravity and irradiation effects are included. We find that most if not all AGN discs are unstable. Observed AGN therefore represent the outburst state, although some or all quasars could constitute a steady population having markedly higher fuelling rates than other AGN. It has important implications for the AGN mass supply and for the presence of supermassive black holes in nearby spirals. ", "introduction": "Accretion discs are a nearly ubiquitous feature of close binary systems, and their presence is widely invoked in models of active galactic nuclei (AGN). A major feature of the discs in binaries is the thermal--viscous instability driven by hydrogen ionization (Meyer \\& Meyer--Hofmeister 1982, Smak 1982). It is now commonly accepted that this instability drives outbursts in cataclysmic variables and soft X--ray transients. Lin \\& Shields (1986) showed by a local stability analysis that this instability can also operate in accretion discs thought to be present around supermassive black holes in AGN. They concluded that these discs were unstable at radii ($ \\approx 10^{15} - 10^{16}$ cm), where the surface temperature is several thousand degrees. The expected characteristic time scale for this instability is $10^4 - 10^7$ years. Because of its generic nature, the ionization instability plays a dominant role in characterizing the observed behaviour of the host systems. In the binary context, attempts to understand the precise conditions (mass of accreting object, accretion rate) under which it occurs have been at least partially successful (e.g. Smak, 1982, van Paradijs 1996; King, Kolb, \\& Burderi, 1996; King, Kolb \\& Szuszkiewicz, 1997, and references therein). These studies show that self--irradiation of the disc by the central X--ray source has a determining effect on the disc stability if the accreting object is compact, as in soft X--ray transients (see below). Delineating the stable and unstable disc regions is equally important for AGN. If the instability is present in AGN discs the suppression of central accretion in the quiescent state means that we can identify only the outburst states of unstable systems as AGN. Two important consequences follow: (1) quiescent AGN must appear as quite normal galaxies, and (2) the average mass fuelling rate in many if not all AGN is much lower than implied by their current luminosities. This in turn limits the masses that their central black holes are expected to reach. The idea of intermittent activity in AGN was already suggested by Shields \\& Wheeler (1978). They noticed that the fuelling problem could be solved if active nuclei store mass during quiescence and this mass then feeds the hole for a shorter period of intense activity. The thermal--viscous hydrogen ionization instability found to operate in AGN accretion discs (Lin \\& Shields, 1986) is capable of triggering such behaviour. Clarke \\& Shields (1989), Mineshige \\& Shields (1990), Cannizzo \\& Reiff, 1992, Cannizzo (1992) studied the full range of black hole masses and accretion rates in order to determine the observational consequence of the instability for the AGN population. Siemiginowska \\& Elvis (1997) attempted to reproduce the observed luminosity function, assuming that this mechanism operates in all AGN. Our aim here is to decide if the ionization instability still operates in AGN when irradiation effects are included. As we have seen irradiation is central to the discussion of disc stability in soft X--ray transients. Further, irradiation is often thought to dominate the disc emission (e.g. Collin--Souffrin, 1994). For both reasons it is vital to include it in any attempt to decide the disc stability. The actual form of the instability when irradiation is included is outside the scope of our paper. Siemiginowska, Czerny \\& Kostyunin (1996) have performed studies for particular black hole masses and accretion rates, with assumed forms of irradiation. There is a simple criterion for the instability to appear: the disc must contain regions with effective temperature $T_{\\rm eff}$ close to the value $T_{\\rm H}$ at which hydrogen is ionized. In practice $T_{\\rm H}$ depends on the density and may be quite different in different environments; we shall consider a range of values in this paper. However the criterion is not easy to use in this form, as one does not in general know the radial distribution of the accretion rate, and thus the run of $T_{\\rm eff}$, in a time--varying disc. Accordingly one usually uses the criterion in an indirect form: a disc with a given constant accretion rate $\\dot M$ is self--consistently steady only if $T_{\\rm eff} > T_{\\rm H}$ throughout it. If the criterion fails we may expect outbursts, although the precise nature of these will depend for example on the detailed behaviour of the disc viscosity. This version of the stability criterion is easy to apply. Since $T_{\\rm eff}$ always decreases with disc radius $R$ in a steady disc, the condition is most stringent at the outer disc radius $R_{\\rm out}$, so we need apply it only there. If the disc's only source of energy is local viscous dissipation we have \\begin{equation} \\left[ T_{\\rm eff}(R) \\right] ^4 = {3GM\\dot{M}\\over 8\\pi R^3\\sigma}f , \\label{eqa} \\end{equation} (e.g. Frank, King \\& Raine (1992); all symbols are explained after equations (2--9)), and the criterion is simply $T_{\\rm eff}(R_{\\rm out}) > T_{\\rm H}$. In a binary system we can estimate $R_{\\rm out}$ with reasonable accuracy as 70\\% of the Roche lobe radius of the accreting star, and the problem is now well determined. Using this approach, Smak (1982) successfully divided outbursting cataclysmic variables (dwarf novae) from the persistent systems (novalikes). The extension to low--mass X--ray binaries is complicated by the fact that the dominant heat source for the disc is not local viscous dissipation (equation (\\ref{eqa})), but irradiation by the central X--rays. The instability is similarly suppressed if the disc surface temperature given by irradiation exceeds $T_{\\rm H}$ (Tuchman, Mineshige \\& Wheeler, 1990). Provided that due account is taken of this, one can again successfully divide the outbursting systems (soft X--ray transients) from the persistent systems (van Paradijs, 1996; King, Kolb \\& Szuszkiewicz, 1997). The key feature, as in the unirradiated case, is that the edge temperature of the disc can be simply expressed in terms of $\\dot M$ and $R_{\\rm out}$, without any need to solve for the full internal disc structure. In both the CV and LMXB cases there are important consequences for the study of the binary evolution (e.g. King, Kolb \\& Burderi, 1996), which gives a connection between $\\dot M, M$ and $R_{\\rm out}$. The extension of this approach to AGN is more complicated; here the outer edge of the disc is no longer determined by the simple Roche lobe condition which holds in binaries, but by the requirement that the disc becomes locally self--gravitating (see equation (\\ref{13}) below). This condition requires a knowledge of the disc density at the outer edge, so we are now required to solve the full global structure of the steady disc to find $R_{\\rm out}$. Thus we examine all possible, stationary, optically thick, geometrically thin disc models relevant for AGN. If these correspond to stable states, AGN discs may be globally steady, and require fuelling at the currently inferred central accretion rates. If not, they will be the outburst states, and the required fuelling rates will be lower than the current central accretion rate. ", "conclusions": "Our aim in this study was to investigate whether the thermal--viscous ionization instability operates in AGN in the presence of irradiation. We have studied stationary, optically thick, geometrically thin discs, in the range of accretion rates and central black hole masses for which these models are self--consistent. It is worth mentioning here that advection dominated optically thin discs can in principle coexist in some particular regions of the parameter space, but which type of the solution will be actually chosen in nature is still an open question. We used a very simple analytic criterion to determine the stability of each model; if the disc is hot enough for hydrogen to be completely ionized everywhere all way out till its self--gravity radius the ionization instability cannot operate. We identify such hot regions in the relevant parts of $\\dot{M}$ -- $M$ plane and show them as grey (for $\\alpha$--discs) and hatched (for $\\beta$--discs) areas in Figure 2. Unlike other authors (Clarke \\& Shields, 1989; Mineshige \\& Shields, 1990, Cannizzo, 1992) we consider only the upper stable branch of the whole cycle, where our method is appropriate. A major advantage of our approach is that we do not need a complicated discussion of the limit cycle. This method proved successful in similar studies of accretion discs in X--ray binaries. We gave careful consideration to the opacities used in our calculations. There are only small differences between results using opacities from Mazzitelli (1989) and Cox \\& Stewart (1970). However differences appear when using simple fitting formulae such as (10) instead of (11) (see Figure 1): it is important to check carefully that a particular fit found in the literature is appropriate for the range of temperatures and densities used in a given problem. Another result of our study is that for $\\alpha \\gta 0.003$ the region between region $a$ and $c$ differs from the standard Shakura--Sunyaev region $b$. We denote it $b^*$. It is radiation pressure dominated, but the main source of opacity is true absorption. We have confirmed the existence of this region in numerical calculations of global disc structure performed using the Cox \\& Stewart (1970) opacity tables. It is interesting that $b^*$ is stable against disc instabilities triggered by radiation pressure (Pringle, 1976): while irradiated it might significantly change its properties. In Figure 3 we compare our results with those based on detailed studies of the outburst cycle over the parameter space considered by various authors. The dotted lines are from Mineshige \\& Shields (1990), dotted--dashed from Clarke \\& Shields (1989), long--dashed from Cannizzo (1992) and the bold lines from this paper. The short--dashed line gives the Eddington limit. Our results for non--irradiated disc are in good agreement with those obtained previously. Our main result, quite contrary to the case of close binaries, is that irradiation does not change the borders between unstable and stable (partially or completely ionized) regions. In other words, irradiation by a central point source is unable to stabilize the whole disc out to its self--gravity radius. An important reason for this is that one of the effects of such irradiation is to move the self--gravity radius even farther out from the central black hole. The irradiated disc structure for low--luminosity, low--mass objects differs from that of the equivalent discs without irradiation (regions C$^+$ in Figure 2). Thus the actual appearance of the ionization instability might well be affected. This can be studied only by detailed calculations of thermal limit cycles in the presence of irradiation. We see from Figure 2 that in general AGN discs will be subject to the ionization instability, even if they are irradiated by a central point source. For typical AGN luminosities, corresponding to central accretion rates $\\lta 10^{-2}M_{\\odot}$~yr$^{-1}$, we see from Figure 2 that it is inconsistent to assume that the disc is stable. Since central accretion (and thus e.g. X--ray emission) is suppressed in the quiescent state, all observed AGN must presumably be identified as such only in their outburst states (which last $\\gta 10^{3}$ yr). Thus AGN currently observed to have central accretion rates below the stability limits $\\sim 10^{-1} - 10^{-2}M_{\\odot}$ yr$^{-1}$ shown in Figure 2 must actually have considerably lower fuelling rates. Even rather brighter observed AGN need not be steady systems, but may simply represent the outburst states of unstable disc with fuelling rates below the stability limits. As pointed out in the Introduction, if most AGN discs are unstable, then in quiescence these systems must be indistinguishable from normal galaxies. Moreover the mass fuelling rates needed to power AGN must be much lower than implied by their current luminosities. If the duty cycle for the outburst can be made short enough ($\\lta 10^{-2}$), no fuelling rates greater than about $10^{-2}\\ M_{\\odot}$~yr$^{-1}$ would be needed in AGN. This would also remove the problem that the remnant black holes are predicted to have excessively high masses if accretion is continuous (Cavaliere \\& Padovani 1988). Alternatively, since most quasars have observed central accretion rates above the stability limits in Figure 2, some or all of them could have steady discs. This group would then form a separate class with much higher fuelling rates $\\dot M \\sim 0.1 - 1\\ M_{\\odot}$~yr$^{-1}$. It is not easy to decide between these possibilities by looking at detailed properties of the individual systems, as outbursting discs rapidly take on a quasi--steady surface density profile (cf Cannizzo, 1993: this property is well known in the context of cataclysmic variables, where the persistent systems -- novalike variables -- look like dwarf novae in permanent outburst). A complicating feature is that many of the objects with high steady fuelling rates would be subject to the radiation--pressure (Lightman--Eardley) instability. We conclude that many (if not all) AGN represent the outburst state of a thermal--viscous disc instability. We should then consider candidates for the quiescent state. It is tempting to suggest that this may comprise most or all ``normal'' spirals. Galaxies such as our own could therefore harbour moderately massive ($10^6 -10^8M_{\\odot}$) black holes in their nuclei. {\\bf Acknowledgements} We thank Ulrich Kolb for valuable discussions. This work is supported by a PPARC Rolling Grant for theoretical astrophysics to the University of Leicester. ARK gratefully acknowledges the support of a PPARC Senior Fellowship. \\clearpage" }, "9803/astro-ph9803131_arXiv.txt": { "abstract": "The Solar Diameter Monitor measured the duration of solar meridian transits during the 6 years 1981 to 1987, spanning the declining half of solar cycle 21. We have combined these photoelectric measurements with models of the solar limb-darkening function, deriving a mean value for the solar near-equatorial radius of 695.508 $\\pm$ .026 Mm. Annual averages of the radius are identical within the measurement error of $\\pm$ .037 Mm. ", "introduction": "The Sun is the only star for which reasonably precise values of the mass, surface radius and luminosity are known. The solar mass $\\Msun$ is known from planetary motion, with accuracy limited only by the uncertainty in the gravitational constant $G$. The solar radius can in principle be obtained from direct optical measurement of the solar angular diameter, given the very accurate determinations of the mean distance between the Earth and the Sun. In solar modeling, the value $\\Rsun = 695.99 \\Mm$ (Allen 1973) has been commonly used. The models are calibrated to this photospheric radius, in the present paper defined by the point in the atmosphere where the temperature equals the effective temperature, by adjusting some measure of the convective efficacy, such as the mixing length. Recent accurate observations of solar f-mode frequencies from the SOI/MDI instrument on the SOHO satellite (e.g. Kosovichev {\\etal} 1997) have raised some doubts over this value of $\\Rsun$. The frequencies of these modes are predominantly determined by $G \\Msun/\\Rsun^3$. By comparing the observed frequencies with frequencies of solar models calibrated to $\\Rsun = 695.99 \\Mm$ Schou {\\etal} (1997) and Antia (1998) concluded that the actual solar radius was smaller by about $0.3 \\Mm$ than the assumed radius of the model. Other aspects of the modeling of the solar f modes may affect their frequencies at this level (e.g. Campbell \\& Roberts 1989; Murawski \\& Roberts 1993; Ghosh, Antia \\& Chitre 1995). Thus it is obviously important to obtain independent verification of the proposed correction to the solar radius. There are indeed significant uncertainties associated with the currently adopted radius value. These are related to the problem of the definition of the solar limb adopted in the radius determinations, and the reduction of the measured value to the photosphere. It is not clear how the value quoted by Allen (1973) was obtained. However, it appears that the more recent determinations, which are generally consistent with Allen, in most cases refer to the inflection point of the solar limb intensity. According to solar atmospheric models this corresponds to a height of about $0.3 \\Mm$ above the photosphere, thus perhaps accounting for the radius correction inferred from the f-mode frequencies. The uncertainty in the precise definition of the measured values of the solar radius highlights the need to combine the observations with careful modeling of the quantity that is observed. Here we consider a long series of observations obtained with the High Altitude Observatory's Solar Diameter Monitor (Brown {\\etal} 1982). This is based on a definition of the solar limb which minimizes the effect of seeing (Hill, Stebbins \\& Oleson 1975). By combining daily data obtained over more than 6 years, extending between solar maximum and solar minimum, the possible effects of solar activity can be checked. The analysis of the data is carried out by means of a model of the solar limb intensity, following as closely as possible the actual procedure used in the reduction of the data and testing for the effects of seeing. In this way we have eliminated several of the uncertainties affecting earlier determinations to arrive at what we believe to be an accurate measure of the solar photospheric radius. ", "conclusions": "We adopted the modified IAU (1976) value of 1.4959787066 $\\times 10^5$ Mm (Astronomical Almanac, 1997) for the astronomical unit, and adjusted this value by -4.678 Mm to account for the mean displacement between the telescope's noontime location and the Earth's center, and by +0.449 Mm for the displacement of the Sun's center relative to the barycenter of the Earth-Sun system. This distance, combined with $D_0$ from Eq. (2), yields the Sun's apparent radius. Applying the model corrections described in the last section, we obtain $$ \\Rsun = (695.5260 \\pm 0.0065 ) \\Mm\\qquad \\hbox{\\rm for Model 1} $$ $$ \\Rsun = (695.4892 \\pm 0.0065 ) \\Mm\\qquad \\hbox{\\rm for Model 2} $$ We estimate the modeling errors to be $1/\\sqrt 2$ of the difference between these estimates, or about 0.020 Mm. Based on the uncertainties in the geometric corrections that were made to the measured radius, we estimate the systematic errors in the measured value to be 0.015 Mm, or about twice as large as the random errors. Averaging our results for Models 1 and 2, and adding the various error sources in quadrature, we arrive at our final estimate of $$ \\Rsun = (695.508 \\pm 0.026 ) \\Mm $$ The inferred solar photospheric radius is smaller by about $0.5 \\Mm$ than the normally used value of $695.99 \\Mm$ (Allen 1976). A review of recent observations was given by Schou {\\etal} (1997), concluding that these were consistent with an angular diameter of $1919.26\\arcsec \\pm 0.2\\arcsec$, corresponding to Allen's value of $\\Rsun$. This is also consistent with the observed value obtained here (cf. eq. 2). However, it appears that the observations considered by Schou {\\etal} refer to the inflection point of intensity (or, in one case, to an FFTD determination) and hence do not contain the correction to photospheric radius. Such a correction, taking into account the observational characteristics, is an essential part of the radius determination. Some confirmation of the reliability of the modeling comes from the comparison in Fig. 2 of computed and observed slopes of the limb position as function of the scan widths. Nevertheless, it is striking that, as indicated by the difference between Models 1 and 2, the major uncertainty in $\\Rsun$ appears to come from the modeling. Indeed, it is evident that the real solar atmosphere is substantially more complicated than the one-dimensional model resulting from the ATLAS code or the mean model obtained from the hydrodynamical simulations. A more accurate determination of the radius correction can probably be obtained from a detailed calculation of the limb intensity, taking into account the inhomogeneous nature of the relevant layers, on the basis of the simulations. Such an investigation is beyond the scope of the present paper, however. We find no significant variation in the observed diameter during the observation period (cf. Fig. 2); annual averages of the radius for the years 1981 to 1987 all agree within their measurement errors of $\\pm .037$ Mm. These limits are substantially smaller than diameter changes reported previously for the same interval of time (e.g. Ulrich \\& Bertello 1995, Laclare {\\it et al.} 1996), but are in agreement with measurements by Wittman (1997). On the other hand, the limb-position slope shows fairly substantial variations. We also note that during solar maximum, the daily slope values tended to be highly variable as well as small in magnitude; this suggests that the long-term variation may result from localized activity-dependent features such as faculae. It is plausible that the previously inferred variations in solar diameter with solar activity is in fact a reflection of such variations in the limb-darkening slope. It is interesting that the value of $\\Rsun$ obtained here is somewhat smaller than that inferred from the solar f-mode frequencies, indicating additional contributions to the differences between the observed and model values of these frequencies. This issue, and the effects of the reduction of the model radius on the helioseismically determined structure of the solar interior will be considered elsewhere. We note, however, that Antia (1998) and Schou {\\etal} (1997) found significant effects on the helioseismically inferred sound speed from corresponding radius changes." }, "9803/astro-ph9803307_arXiv.txt": { "abstract": "I review the status of observational determinations of central masses in nearby galactic nuclei. Results from a variety of techniques are summarized, including ground-based and space-based optical spectroscopy, radio VLBI measurements of luminous water vapor masers, and variability monitoring studies of active galactic nuclei. I will also discuss recent X-ray observations that indicate relativistic motions arising from the accretion disks of active nuclei. The existing evidence suggests that supermassive black holes are an integral component of galactic structure, at least in elliptical and bulge-dominated galaxies. The black hole mass appears to be correlated with the mass of the spheroidal component of the host galaxy. This finding may have important implications for many astrophysical issues. ", "introduction": "The discovery of quasars in the early 1960's quickly spurred the idea that these amazingly powerful sources derive their energy from accretion of matter onto a compact, extremely massive object, most likely a supermassive black hole (SMBH; Zel'dovich \\& Novikov 1964; Salpeter 1964; Lynden-Bell 1969) with $M\\,\\approx\\,10^6-10^9$ \\solmass. Since then this model has provided a highly useful framework for the study of quasars, or more generally, of the active galactic nucleus (AGN) phenomenon (Rees 1984; Blandford \\& Rees 1992). Yet, despite its success, there is little empirical basis for believing that this model is correct. As pointed out by Kormendy \\& Richstone (1995, hereafter KR), our confidence that SMBHs must power AGNs largely rests on the implausibility of alternative explanations. To be sure, a number of characteristics of AGNs indicate that the central engine must be tiny and that relativistic motions are present. These include rapid X-ray variability, VLBI radio cores, and superluminal motion. However, solid evidence for the existence of SMBHs in the centers of galaxies has, until quite recently, been lacking. As demonstrated by Soltan (1982), simple considerations of the quasar number counts and standard assumptions about the efficiency of energy generation by accretion allows one to estimate the mean mass density of SMBHs in the universe. The updated analysis of Chokshi \\& Turner (1992) finds $\\rho_{\\bullet}\\,\\approx\\,2 \\times 10^5 \\epsilon_{0.1}^{-1}$ \\solmass\\ Mpc$^{-3}$ for a radiative efficiency of $\\epsilon\\,=\\,0.1\\epsilon_{0.1}$. Comparison of $\\rho_{\\bullet}$ with the $B$-band galaxy luminosity density of 1.4\\e{8}$h$ \\solum\\ Mpc$^{-3}$ (Lin \\etal 1996), where the Hubble constant $H_{\\rm 0}$ = 100$h$ \\kms\\ Mpc$^{-1}$, implies an average SMBH mass per unit stellar luminosity of $\\sim$1.4\\e{-3}$\\epsilon_{0.1}^{-1} h^{-1}$ \\solmass/\\solum. A typical bright galaxy with $L_B^*\\,\\approx\\,10^{10} h^{-2}$ \\solum\\ potentially harbors a SMBH with a mass \\gax $10^7\\epsilon_{0.1}^{-1}h^{-3}$ \\solmass. These very general arguments lead one to conclude that ``dead'' quasars ought to be lurking in the centers of many nearby luminous galaxies. The hunt for SMBHs has been frustrated by two principal limitations. The more obvious of these can be easily appreciated by nothing that the ``sphere of influence'' of the hole extends to $r_{\\rm h}\\,\\simeq\\,G M_{\\bullet}/ \\sigma^2$ (Peebles 1972; Bahcall \\& Wolf 1976), where $G$ is the gravitational constant and $\\sigma$ is the velocity dispersion of the stars in the bulge, or, for a distance of $D$, $\\sim$1\\asec ($M_{\\bullet}/2\\times10^8$ \\solmass)($\\sigma$/200 \\kms)$^{-2}$($D$/5 Mpc). Typical ground-based observations are therefore severely hampered by atmospheric seeing, and only the heftiest dark masses in the closest galaxies can be detected. The situation in the last few years has improved dramatically with the advent of the {\\it Hubble Space Telescope (HST)} and radio VLBI techniques. The more subtle complication involves the actual modeling of the stellar kinematics data, and in this area much progress has also been made recently as well. Here I will highlight some of the observational efforts during the past two decades in searching for SMBHs, concentrating on the recent advances. Since this contribution is the only one that discusses nuclear BHs aside from that in the Milky Way (Ozernoy, these proceedings) and in NGC 4258 (Miyoshi, these proceedings), I will attempt to be as comprehensive as possible, although no claim to completeness is made, as this is a vast subject and progress is being made at a dizzying pace. To fill in the gaps, I refer the reader to several other recent review papers, each of which has a slightly different emphasis (KR; Rees 1998; Richstone 1998; Ford \\etal 1998; van der Marel 1998). ", "conclusions": "" }, "9803/astro-ph9803288_arXiv.txt": { "abstract": "We consider the formation of low--mass X--ray binaries containing accreting neutron stars via the helium--star supernova channel. The predicted relative number of short--period transients provides a sensitive test of the input physics in this process. We investigate the effect of varying mean kick velocities, orbital angular momentum loss efficiencies, and common envelope ejection efficiencies on the subpopulation of short--period systems, both transient and persistent. Guided by the thermal--viscous disk instability model in irradiation--dominated disks, we posit that short--period transients have donors close to the end of core--hydrogen burning. We find that with increasing mean kick velocity the overall short-period fraction, $s$, grows, while the fraction, $r$, of systems with evolved donors among short-period systems drops. This effect, acting in opposite directions on these two fractions, allows us to constrain models of LMXB formation through comparison with observational estimates of $s$ and $r$. Without fine tuning or extreme assumptions about evolutionary parameters, consistency between models and current observations is achieved for a regime of intermediate average kick magnitudes of about 100--200\\,km\\,s$^{-1}$, provided that (i)~orbital braking for systems with donor masses in the range $1-1.5\\,\\msun$ is weak, i.e., much less effective than a simple extrapolation of standard magnetic braking beyond $1.0\\,\\msun$ would suggest, and (ii)~the efficiency of common envelope ejection is low. ", "introduction": "The structure and properties of accretion disks around non-magnetic compact objects in close binaries are primarily determined by the rate at which matter is supplied by the Roche--lobe filling donor star. If this rate is smaller than a critical value $\\dmc$, the disk is subject to a thermal--viscous instability and undergoes a limit cycle evolution alternating between a hot and a cool state, the outburst and quiescent phases. This model has been successfully applied to dwarf nova outbursts in cataclysmic variables (CVs), where the accretor is a white dwarf (WD) (for reviews see, e.g., Cannizzo\\markcite{C93} 1993; Osaki\\markcite{O96} 1996), and also to X--ray transient outbursts in low-mass X-ray binaries (LMXBs), where the accretor is a neutron star or a black hole (for reviews see, e.g., Lasota\\markcite{L96} 1996; Wheeler\\markcite{W98} 1998). In the case of LMXBs, where the accretion efficiency and hence the X-ray luminosity from the compact object is higher than in CVs, there is strong evidence that self--irradiation is important (see, e.g., van~Paradijs \\& McClintock\\markcite{vP94} 1994; King \\& Ritter\\markcite{KR98} 1998). X--ray irradiation from the central accretor raises the disk temperature for a given mass transfer rate, thereby suppressing the instability. As a consequence, $\\dmc$ is significantly smaller in LMXBs than in CVs, consistent with observations (van~Paradijs\\markcite{vP96} 1996). The strength of the irradiation in LMXBs depends on the nature of the compact object. In the neutron--star case, the irradiating source is equivalent to a point source at the center of the disk, while, in the black--hole case, the lack of a hard stellar surface implies that the irradiating source is only the innermost disk, which is weaker by a factor about equal to the relative disk thickness, the other system parameters being constant (King, Kolb, \\& Szuskiewicz\\markcite{KK97} 1997). Disk irradiation has important consequences for short--period LMXBs. These systems evolve towards shorter orbital periods $P$ under the influence of orbital angular momentum losses $\\dot J$ (henceforth ``j--driven'' systems), caused by a magnetic stellar wind from the donor star (magnetic braking) or by gravitational radiation. Assuming that $\\dot{J}_{\\rm MB}$ from magnetic braking (MB) is independent of the nature of the accretor, the same orbital braking formalism must apply for both CVs and LMXBs. For a given MB law, i.e., for a given dependence of $\\dot J$ on binary parameters, the mass transfer rate (proportional to the fractional angular momentum loss rate) is smaller in black--hole binaries than in neutron--star binaries, as the total angular momentum increases with primary mass. This, together with the higher values for $\\dmc$, leads to transient behavior for all j--driven black--hole binaries, as observed. The converse appears to be true for neutron--star systems. The MB transfer rate may be up to two orders of magnitude larger than $\\dmc$ (King, Kolb, \\& Burderi\\markcite{KK96} 1996), suggesting that there are no or only very few neutron star transients. However, it is clear from observations that the fraction of transients among short--period neutron--star LMXBs is non--negligible. Five out of 23 LMXBs with a confirmed or possible neutron star primary and with periods in the range $3\\,{\\rm h} < P < 20\\,{\\rm h}$ are classified as transient (Ritter \\& Kolb\\markcite{R98} 1998). Although it is difficult to estimate the {\\em intrinsic} value of this fraction, it may be even larger than $5/23$ (see \\S\\,5.1 below). A possible resolution of this apparent contradiction between the disk-instability model and the observations has been suggested by King, Kolb, \\& Burderi\\markcite{KK96} (1996) and King \\& Kolb\\markcite{K97} (1997, hereafter KK97). The mass transfer rates in j--driven LMXBs\\footnote{Hereafter, we will refer to neutron--star LMXBs simply as LMXBs, unless otherwise stated.} with somewhat evolved donor stars (close to the end of core--hydrogen burning), are significantly smaller than in systems with unevolved donors, therefore favoring transient behavior. Then the large observed transient fraction demands that the contribution of these evolved systems to the total population is correspondingly large. Under the assumption of spherically symmetric supernovae (SN) and for strong magnetic braking at high ($\\gtrsim 1.2\\,{\\rm M}_\\odot$) donor star masses, KK97 showed that LMXBs forming via the standard evolutionary channel involving a helium star supernova (Sutantyo\\markcite{S75} 1975; van den Heuvel\\markcite{vdH83} 1983) do indeed have this property. However, for weaker magnetic braking (constrained by rotational velocity data for F stars), Kalogera \\& Webbink\\markcite{K98} (1998; hereafter KW98) showed that j--driven LMXBs form only if supernovae are asymmetric. KK97 also showed, in a qualitative way, that asymmetric SNe with on average large kick velocities imparted to the neutron stars would inevitably lead to a large number of unevolved systems in the population, and hence a small number of transient systems among the short-period LMXB population, contrary to observations. This qualitative conclusion by KK97 appears to be in contradiction to numerous pieces of evidence for the existence of rather substantial kick velocities (Kaspi et al.\\ \\markcite{Kp96}1996; Hansen \\& Phinney\\markcite{H97} 1997; Lorimer, Bailes, \\& Harrison\\markcite{L97} 1997; Fryer \\& Kalogera\\markcite{F97} 1997; Fryer, Burrows, \\& Benz\\markcite{F98} 1998), so the need for a quantitative study of the problem arises. The formation of LMXBs including the effect of natal neutron--star kicks has been studied in detail by KW98. Here, we use these detailed synthesis models to address the issue of the transient fraction among LMXBs and investigate its dependence on different evolutionary parameters. The increase of this fraction in the model populations is accompanied by an increase of the fraction of systems with donors that have evolved beyond the main sequence, i.e., of long--period systems driven by nuclear expansion of the secondary (``n--driven'' systems). Our goal is twofold: first, to disentangle the differential dependences on model parameters of the two observable quantities, the predicted fraction of ``j--driven'' systems among all LMXBs and the predicted transient fraction among ``j--driven'' LMXBs. Second, to constrain quantitatively the input parameters by requiring that both predicted fractions are consistent with current observations. These two observational constraints operate in opposite directions, and this allows us to derive limits on the mean magnitude of natal kicks imparted to neutron stars, the strength of magnetic braking, and the efficiency with which orbital energy is consumed during the common envelope phase prior to the supernova. As the observational sample increases in the future, the model calculations presented here can be used to tighten these constraints still further. In \\S\\,2, we review the evolutionary picture of j--driven LMXBs and derive a simplified criterion for transient behavior. In \\S\\,3, we describe the different models for LMXB formation considered here, while the results and basic effects of varying model parameters are presented in \\S\\,4. In \\S\\,5, we first (\\S\\,5.1) describe the observed sample and derive the observational constraints, and then (\\S\\,5.2) evaluate the models based on these constraints. We conclude in \\S\\,6 with a discussion of our results and of possible extensions of the present study. ", "conclusions": "We applied the population synthesis techniques developed by KW98 to investigate the influence of varying mean supernova--kick magnitudes, orbital angular momentum loss strengths, and common envelope efficiencies on the population of transient and persistent short--period LMXBs that form via the helium star supernova channel. A main premise of our study is the identification of short--period transients with systems where the donor is significantly nuclear--evolved, close to the end of core--hydrogen burning. We tested our models against two properties inferred from the observed sample: that a significant fraction ($s \\ga 50\\%$) of nascent LMXBs are short--period, and that a significant fraction of these ($r \\ga 20\\%$) have donors close to the end of core--hydrogen burning. We found that any combination of model parameters that results in a large fraction of short-period systems with evolved donors, hence high values of $r$, at the same time results in the formation of many systems with donors that have evolved beyond the main sequence, and hence leads to low values of $s$. It is exactly these two counteracting effects that allow us to constrain the three model parameters. With an increasing mean kick magnitude $s$ grows, while $r$ drops. Model predictions for $s$ and $r$ are consistent with observational estimates in an intermediate regime of moderate mean kick magnitudes, $\\vmean \\simeq 100-200$~km/s, if (i)~the orbital braking for systems with donor masses $1\\la M_d \\la 1.5\\,\\msun$ is weak, i.e., much less effective than a simple extrapolation of magnetic braking beyond $1\\,\\msun$ would suggest, and (ii)~the efficiency of common envelope ejection is low ($\\alpha_{\\rm CE} \\la 0.5$). Consistency with observational estimates could also be achieved in the absence of kicks or for a very large mean kick velocity, but {\\em only} in combination with a fine-tuned common-envelope efficiency and/or fairly extreme assumptions on the efficiency of magnetic braking at masses $\\ga 1\\,\\msun$ (very high for small kicks, almost negligible for large kicks). However, in view of the various uncertainties that enter our study we cannot unambiguously rule out these models. Overall, we have shown that the model predictions are sufficiently sensitive to the variation of the three input parameters to place constraints on them when comparing to observations. Given accurate input from the observational sample, the constraints can be reasonably tight (for example, had the adopted values of $s$ and $r$ been known within a factor of 2, we could unambiguously exclude a single peak of kick magnitudes at $\\lesssim 50$ or $\\gtrsim 300$\\,km\\,s$^{-1}$). Nevertheless, significant uncertainties exist that have their origin mainly in two areas: secular evolution of j--driven LMXBs and observational selection effects, which prevent us from accurately inferring the intrinsic properties of the LMXB population from the observed sample. In the first area, a systematic study of j--driven LMXB evolution with detailed stellar models is needed to relate transient behavior quantitatively to the degree of evolution of the donor for different orbital angular momentum loss rates. This will also allow one to examine in detail the transition regime between j--driven and n--driven systems and to quantify the relative lifetimes of persistent and transient systems. The efficiency of magnetic braking affects both the lifetime of the systems and the critical degree of evolution for transient behavior. An increasing braking strength restricts transient behavior to donors closer to the end of core--hydrogen burning. At the same time, it leads to higher mass transfer rates in persistent systems, and hence shorter relative lifetimes. The first effect increases the critical degree of evolution, $f_0$, whereas the second reduces the intrinsic transient fraction, $r_0$, derived from the observed sample. Therefore, the consistency criterion, $\\rf > r_0$, moves along a typical curve $\\rf$ (Fig.\\,\\ref{r1} and \\ref{r2}), so that the choice of successful or unsuccessful models is not, to zeroth order, affected by the neglect of these dependences. We note that another long--standing problem in the evolution of LMXBs, the fate of j--driven systems once they approach to orbital periods of 3\\,h, and the apparent lack of systems with $P \\la 3$\\,h (see KK97 for a discussion), does not affect our study as long as any comparison is restricted to systems with periods longer than 3\\,h. The Skumanich--type magnetic braking formulation adopted here is by no means the only possible parametrization of the orbital angular momentum losses. Constraints on the real functional form could come from further analysis of stellar rotational rates in open clusters of different age. Some studies (e.g., Krishnamurthi et al.\\markcite{K97} 1997, and references therein) suggest that $\\dot J$ becomes a less steep function of $\\omega$ above a certain critical angular velocity. Repeating our model calculations with such more general magnetic braking laws seems worthwhile only once the critical degree of evolution, $f_{\\rm 0}$, for transient behavior can be estimated more quantitatively. Progress in the second area of uncertainty requires a systematic study of observational selection effects on X--ray binaries (given the X-ray instruments used so far). This is hampered by the difficulty of calculating the spectral distribution of the emergent X--ray flux in a system with given period and transfer rate. The assessment of the relative completeness of the different subgroups involved (black--hole, neutron--star, transient, persistent, short--period, and long--period systems) will certainly gain reliability when future observations increase the known sample and the number of systems with determined binary and transient parameters. The completeness is further affected by a possible dependence of the transient outburst recurrence time on orbital period. Giant donor systems have larger disks, which take longer to reestablish the critical pre--outburst surface density. If the recurrence time is systematically longer in long--period systems, the short--period fraction could be significantly smaller than estimated. This in turn would make models with even smaller mean kick velocities acceptable. Despite these uncertainties, the preference for moderate mean kick velocities of $\\simeq 100-200$~km/s inferred from our study is likely to persist even for a more detailed treatment of secular evolution and a better understanding of selection effects. This preference seems to be in conflict with the fairly large {\\em mean} natal kicks traditionally deduced from pulsar proper motions (e.g., $\\vmean \\simeq 500$~km/s is favored by Lorimer et al.\\ 1997). (Note that our study cannot constrain the fraction of very high ($\\gtrsim 500$\\,km\\,s$^{-1}$) kick magnitudes, because this only affects the absolute LMXB birth-rate normalization.) While this might point to an underlying physical difference in the way supernova explosions of type II and type Ib proceed, there is also the possibility that natal pulsar velocities are significantly smaller than this estimate. Indeed, it has been pointed out that a relatively wide variety of qualitatively different distributions are consistent with the observed pulsar velocity distribution (Hansen \\& Phinney 1997; Cordes \\& Chernoff\\markcite{C98} 1998; Fryer et al.\\ 1998; Hartmann\\markcite{H98} 1998). Although our study was limited to Maxwellian kick distributions we can conclude quite generally that the observed short-period transient LMXB population favors kick distributions with a dominant component at moderate mean velocities of about 100 km/s. We conclude by noting that with future progress the observational sample of LMXBs will undoubtedly increase and improve in quality, and more detailed calculations of the secular evolution of LMXBs will become available. The model calculations we have presented here, of the fraction of short-period LMXBs and the transient fraction among them, can then be used to derive even tighter constraints on the supernova and evolutionary parameters." }, "9803/astro-ph9803182_arXiv.txt": { "abstract": "The pulsar PSR B1259$-$63 is in a highly eccentric 3.4-yr orbit with the Be star SS 2883. Timing observations of this pulsar, made over a 7-yr period using the Parkes 64-m radio telescope, cover two periastron passages, in 1990 August and 1994 January. The timing data cannot be fitted by the normal pulsar and Keplerian binary parameters. A timing solution including a (non-precessing) Keplerian orbit and timing noise (represented as a polynomial of fifth order in time) provide a satisfactory fit to the data. However, because the Be star probably has a significant quadrupole moment, we prefer to interpret the data by a combination of timing noise, dominated by a cubic phase term, and $\\dot\\omega$ and $\\dot x$ terms. We show that the $\\dot\\omega$ and $\\dot x$ are likely to be a result of a precessing orbit caused by the quadrupole moment of the tilted companion star. We further rule out a number of possible physical effects which could contribute to the timing data of PSR B1259$-$63 on a measurable level. ", "introduction": "The pulsar PSR B1259$-$63 is a member of a unique binary system. Discovered using the Parkes telescope in a survey of the Galactic plane at 1.5 GHz \\cite{jlm+92a}, it was shown by Johnston et al.~(1992b) to be in a highly eccentric 3.4-yr orbit with a 10th-magnitude Be star, SS 2883. The pulsar period, $P$, is relatively short, 47.8 ms, and the measured period derivative, $\\dot{P}$, gives a pulsar characteristic age, $\\tau_c = P/(2\\dot{P})$, of $3.3 \\times 10^5$ yr and a surface magnetic field of 3.3$\\times 10^{11}$ G. This therefore is a young system, which may evolve through an accretion phase to form a single or binary millisecond pulsar. Rapidly spinning neutron stars can only accrete matter if the co-rotation velocity at the Alfv\\'en radius is less than the Keplerian velocity at the same radius \\cite{bv91}. Equality of these velocities defines the `spin-up line'. At present, PSR B1259$-$63 lies well to the left of the spin-up line, so that accretion onto the neutron star is not possible until either the pulsar slows down or the pulsar magnetic field decays. Timing observations of PSR B1259$-$63, made over a 3.4-yr interval and covering the 1990 August periastron, were reported by Johnston et al.\\ \\shortcite{jml+94}. A phase-connected fit to these data gave parameters for the pulsar and its orbit, and showed that the next periastron would occur on 1994 January 9. This paper also reported optical observations which indicate that the companion star is of spectral type B2e, with a mass of $m_* \\sim 10$ M$_{\\sun}$ and radius $R_* \\sim 6$ R$_{\\sun}$. The break-up velocity at the equator, $v_{{\\rm max}}$, for B2e stars is not very well known; it is estimated to be $\\sim 380$ km s$^{-1}$ by Slettebak et al.\\ (1980) and $\\sim 480$ km s$^{-1}$ by Schmidt-Kaler (1982). Recent work by Porter (1996) suggests that most Be stars rotate at $\\sim 70$ per cent of the break-up velocity. From the mass function, a companion mass of 10 M$_{\\sun}$ and a pulsar mass, $m_p$, of 1.4 M$_{\\sun}$ imply an orbital inclination $i=36\\degr$. The orbital eccentricity is very high, 0.87, and the pulsar approaches within 24 $R_*$ of the companion star at periastron, passing through the circumstellar disk. Extensive observations of the pulsar were made at several radio frequencies before and after the 1994 January periastron, in order to probe the circumstellar environment of SS 2883 (Johnston et al.\\ 1996; Melatos, Johnston \\& Melrose 1995). Observations made between 1990 January and 1994 October were well explained by step changes in the pulsar period at the two periastrons \\cite{mjl+95}, attributed to a propeller-torque spin-down caused by the interaction of the pulsar with the circumstellar matter at the Alfv{\\'e}n radius (Illarionov \\& Sunyaev 1975, King \\& Cominsky 1994, Ghosh 1995). In this paper we report on additional timing observations made using the Parkes radio telescope over the past two years which, together with the earlier data, give a total timing data span of seven years. We find that the timing solution of Manchester et al. (1995) does not fit the recent data and discuss alternative models and their interpretation. ", "conclusions": "At present, the timing observations for the binary pulsar PSR B1259$-$63 span seven years. Because of the gaps in timing observations around the two periastrons and the large timing noise present in this young pulsar, we still are not able to derive a unique timing model to explain the TOAs. Model 1 is a timing solution including a non-precessing Keplerian orbit and timing noise represented as a polynomial of fifth order in time. This model provides a satisfactory fit to the data. The remaining timing residuals are understood as short-term timing noise similiar to that seen in observations of other young pulsars (cf.\\ Foster et al.\\ 1994). Equally good results were obtained by Model 2 and 3. Both timing models contain just a $\\ddot P$ term to account for the long-term behaviour of the timing noise, and $\\dot\\omega$ and $\\dot x$, which both are understood to result from a precession of the orbit. This orbital precession can be explained by the classical spin-orbit coupling caused by the quadrupolar nature of the main-sequence star companion. The corresponding advance of periastron is negative and thus the companion should be tilted by more than $30\\degr$ with respect to the orbital plane (See Fig.\\ 6). This can be explained by a birth kick for the pulsar (cf.\\ Kaspi et al.\\ 1996). Tidal dissipation and frictional drag in the circumstellar matter is shown to be negligible. The influence of the mass loss of the companion is too small to be detectable unless the mass loss is $\\ga 10^{-5} M_\\odot$/yr. At the same time we can exclude a significant orbital period change in the TOAs of PSR B1259--63. PSR B1259$-$63 should show the largest Einstein delay and largest Shapiro delay of all known binary pulsars. The small change in the longitude of periastron makes it impossible to isolate the Einstein delay. The Shapiro delay peaks sharply around periastron and is so far unobservable. Again, we stress that the physical parameters given here for the companion star, $\\theta$, $\\Phi$ and $k$, should be understood as one possible explanation for the significant values of $\\dot\\omega$ and $\\dot x$ in Models 2A and 2B. If the long-term behaviour of the timing noise of PSR B1259$-$63 is not fully modelled by a cubic term ($\\ddot P$), it is possible that rather large fractions of these parameters are not explained by a precession of the orbital plane but have their origin in unmodelled timing noise. The parameters here show that, in principle, all of the $\\dot\\omega$ and $\\dot x$ can arise from classical spin-orbit coupling, for Model 2A in particular. Although Model 1 gives a good fit to the TOAs without making use of the classical spin-orbit coupling, it seems unlikely that the classical spin-orbit coupling is of no importance for this system." }, "9803/astro-ph9803019_arXiv.txt": { "abstract": "We present deep near--infrared images of high redshift radio galaxies (HzRGs) obtained with the Near Infrared Camera (NIRC) on the Keck I telescope. In most cases, the near--IR data sample rest wavelengths free of contamination from strong emission lines and at $\\lambda_{\\rm rest} > 4000$\\AA, where older stellar populations, if present, might dominate the observed flux. At $z > 3$, the rest--frame optical morphologies generally have faint, large--scale ($\\sim$50 kpc) emission surrounding multiple, $\\sim 10$ \\kpc~size components. The brightest of these components are often aligned with the radio structures. These morphologies change dramatically at $2 < z < 3$, where the $K$ images show single, compact structures without bright, radio--aligned features. The linear sizes ($\\sim 10$ \\kpc) and luminosities ($M(B_{\\rm rest}) \\sim -20$ to $-22$) of the {\\it individual} components in the $z > 3$ HzRGs are similar to the {\\it total} sizes and luminosities of normal, radio--quiet, star forming galaxies at $z = 3 - 4$ (Steidel et al.\\ 1996; Lowenthal et al.\\ 1997). For objects where such data are available, our observations show that the line--free, near--IR colors of the $z > 3$ galaxies are very blue, consistent with models in which recent star formation dominates the observed light. Direct, spectroscopic evidence for massive star formation in one of the $z >3$ HzRGs exists (4C41.17, Dey \\etal 1997$a$). Our results suggest that the $z > 3$ HzRGs evolve into much more massive systems than the radio--quiet galaxies and that they are qualitatively consistent with models in which massive galaxies form in hierarchical fashion through the merging of smaller star--forming systems. The presence of relatively luminous sub--components along the radio axes of the $z > 3$ galaxies suggests a causal connection with the AGN. We compare the radio and near--IR sizes as a function of redshift and suggest that this parameter may be a measure of the degree to which the radio sources have induced star formation in the parent objects. We also discuss the Hubble diagram of radio galaxies, the possibility of a radio power dependence in the $K - z$ relation, and its implications for radio galaxy formation. Finally, we present for the first time in published format basic radio and optical information on 3C~257 ($z=2.474$), the highest redshift galaxy in the 3C sample and among the most powerful radio sources known. ", "introduction": " ", "conclusions": "" }, "9803/astro-ph9803255_arXiv.txt": { "abstract": "A linearity test shows $H_0$ to decrease by 7\\% out to $18\\,000\\kms$. The value at $10\\,000\\kms$ is a good approximation to the mean value of $H_0$ over very large scales. The construction of the extragalactic distance scale is discussed. Field galaxies, cluster distances relative to Virgo, and blue supernovae of type Ia yield $H_{0}$\\,(cosmic) with increasing weight; they give consistently $H_{0}=57\\pm7$ (external error). This value is supported by purely physical distance determinations (SZ~effect, gravitational lenses, MWB~fluctuations). Arguments for $H_{0}>70$ are discussed and shown to be flawed. ", "introduction": "\\label{sec:1} The calibration of the cosmic expansion rate $H_0$ consists of two steps. The first step is an investigation of the cosmic expansion field. How {\\em linear\\/} is the expansion? How large are {\\em systematic\\/} deviations from linearity in function of distance? What is the {\\em scatter\\/} of individual objects due to peculiar motions about the mean expansion? Only after these questions are solved can the second step be tackled, i.\\,e. the calibration of the expansion rate in absolute terms. The procedural difference between the two steps is that only redshifts and {\\em relative\\/} distances are needed for an investigation of the characteristics of the expansion field, while the calibration of the present large-scale expansion rate $H_0$ requires in addition the {\\em true\\/} distance of at least one object which demonstratably partakes of the mean expansion. Much confusion about the expansion rate has arisen from equating the velocity-distance ratio of a subjectively chosen object with $H_0$. The determination of $H_0$ from the Virgo, Fornax, or Coma clusters, for instance, is meaningful only if it is demonstrated that they reflect at their moderate distances the mean cosmic expansion. The long-standing problem of correcting the observed mean velocity of the Virgo cluster into the frame of the cosmic expansion field has become a classic (cf. Section~3.2). The Fornax cluster with a velocity of $v\\approx1200\\kms$ cannot be used for the determination of $H_0$, even if a useful distance was known for it, because its unknown peculiar velocity may well be as high as 20\\% of its observed velocity. And the Coma cluster at $v\\approx7000\\kms$, which is sometimes used for the determination of $H_0$, may still have a peculiar velocity component of 10\\%, as the peculiar velocity of $630\\kms$ with respect to the MWB of one other supercluster, i.\\,e. the Local Supercluster, would suggest. The present paper outlines this two-step procedure. In Section~2 the available data are used to map the expansion rate in function of distance well beyond $30\\,000\\kms$, i.\\,e. out to distances where the truly cosmic character of $H_0$ cannot be questioned. Section~3 gives a summary of the various methods of determining distances of field galaxies, of the Virgo cluster, and --- most decisively for $H_0$ --- of distant blue SNe\\,Ia. Methods leading seemingly to $H_{0}>70$ are critically discussed in Section~4. A brief outlook is given in Section~5. ", "conclusions": "\\label{sec:5} A Test for the variation of $H_0$ with distance suggests a decrease by $\\sim\\!7\\%$ from $1000 < v \\le 18\\,000\\kms$. At $v=10\\,000\\kms$ $H_0$ goes through a value close to the mean over very large scales. A system of three interconnected distance scales (field galaxies, cluster distances relative to the Virgo cluster, and most significantly blue SNe\\,Ia) give $H_{0}$ (cosmic) $=57\\pm7$ (external error). Physical distance determinations from the SZ effect, gravitationally lensed quasars, and MWB fluctuations scatter about the same value. A discussion of proposed high values of $H_0$ shows that disagreement focuses on two topics: 1) the true distance of the Virgo cluster, and 2) the appreciation of the Malmquist bias. One may add as item 3) the distance of the E/S0 galaxies in the Fornax cluster; the latter has lower priority because the peculiar motion of this cluster is unknown, and it is poorly tied into the relative distance scale of other clusters. \\bigskip \\noindent {\\small {\\bf Acknowledgement:} Financial support of the Swiss National Science Foundation is gratefully acknowledged. The author thanks his colleagues in the $HST$ team for the luminosity calibration of SNe\\,Ia, i.\\,e. Dres. A.~Sandage, A.~Saha, L.~Labhardt F.\\,D.~Macchetto, and N. Panagia, as well as the many collaborators behind the scenes at the STScI; much of the present understanding of $H_0$ depends on their work. He also thanks Mr.~Bernd Reindl for his excellent help in all computational and technical matters.}" }, "9803/astro-ph9803313_arXiv.txt": { "abstract": "We have obtained medium-resolution spectra of seven UV-bright stars discovered on images of four southern globular clusters obtained with the Ultraviolet Imaging Telescope (UIT). Effective temperatures, surface gravities and helium abundances are derived from LTE and non-LTE model atmosphere fits. Three of the stars have sdO spectra, including M4-Y453 (\\teff\\ = 58800~K, \\logg\\ = 5.15), NGC 6723-III60 (\\teff = 40600~K, \\logg\\ = 4.46) and NGC~6752-B2004 (\\teff\\ = 37000~K, \\logg\\ = 5.25). All seven stars lie along either post-extended horizontal branch (EHB) or post-early AGB evolutionary tracks. The post-early AGB stars show solar helium abundances, while the post-EHB stars are helium deficient, similar to their EHB progenitors. ", "introduction": "Ultraviolet images of globular clusters are often dominated by one or two hot, luminous, ``UV-bright'' stars. The most luminous of these stars are believed to be post-asymptotic giant branch (post-AGB) stars, which go through a luminous UV-bright phase as they leave the AGB and move rapidly across the HR diagram toward their final white dwarf state. Despite their short lifetimes ($\\sim 10^{5}$ yrs), hot post-AGB stars can dominate the total ultraviolet flux of an old stellar population. In particular, hot post-AGB stars probably make a significant (although not the dominant) contribution to the UV-upturn observed in elliptical galaxies (Brown et al.\\ \\cite{brown97}). However, a large uncertainty exists in modeling the contribution of hot post-AGB stars to the integrated spectrum of an old stellar population, due to the strong dependence of the post-AGB luminosity and lifetime on the core mass, which in turn depends on when the stars leave the AGB (Charlot et al.\\ \\cite{charris96}). Also the previous mass loss on the red giant branch (RGB) plays an important r\\^ole here, since it determines the fate of a star during and after the horizontal branch stage: Stars with very low envelope masses settle along the extended horizontal branch (EHB) and evolve from there directly to the white dwarf stage, whereas stars with envelope masses of more than 0.02~\\Msolar will at least partly ascend the AGB. A further uncertainty arises because theoretical post-AGB tracks have been only minimally tested for old, low-mass stars. The last census of hot post-AGB stars in globular clusters was published by de Boer (\\cite{debo87}), but this list is certainly incomplete. The detection of hot post-AGB stars in optical color-magnitude diagrams (CMD's) is limited by selection effects due to crowding in the cluster cores and to the large bolometric corrections for these hot stars. More complete searches are possible for hot post-AGB stars in planetary nebulae, for example, by using \\ion{O}{III} imaging. However, only four planetary nebulae (PNe) were discovered in a recent survey of 133 globular clusters (Jacoby et al.\\ \\cite{jamo97}), of which two were previously known (K648 in M~15, and IRAS 18333-2357 in M~22). Jacoby et al. expected to find 16 planetary nebulae in their sample, on the basis of the planetary nebula luminosity function for metal-poor populations. The origin of this discrepancy is not yet understood, but we mention two possible contributing factors. First, the \\ion{O}{III} search of Jacoby et al. may have missed some old, faint planetary nebulae. Second, Jacoby et al. derive the number of expected PNe from the total cluster luminosity, assuming that all stars in a globular cluster will eventually go through the AGB phase. But in a cluster such as NGC 6752, about 30\\% of the HB population consists of EHB stars (with \\teff\\ $>$ 20,000~K), which are predicted to evolve into white dwarfs without ever passing through the thermally pulsing AGB phase. The exact fraction of stars which follow such evolutionary will depend on the poorly known mass loss rates during the HB and early-AGB phases. While globular clusters with a populous EHB are expected to be deficient in post-AGB stars, they should show a substantial population of less luminous (1.8 $<$ \\logl\\ $<$ 3) UV-bright stars, which can be either post-EHB stars or post-early AGB stars. The population of post-EHB stars is expected to be about 15--20 \\% of the population of EHB stars (Dorman et al. \\cite{dorm93}). The post-early AGB population arises from hot HB stars with sufficient envelope mass to return to the AGB, but which peel off the AGB prior to the thermally pulsing phase (Dorman et al. \\cite{dorm93}). During the two flights of the {\\em ASTRO} observatory in 1990 and 1995, the Ultraviolet Imaging Telescope (UIT, Stecher et al.\\ \\cite{stech97}) was used to obtain ultraviolet ($\\sim 1600$~\\AA) images of 14 globular clusters. The solar-blind detectors on UIT suppress the cool star population, which allows UV-bright stars to be detected into the cluster cores, and the $40'$ field of view of UIT is large enough to image the entire population of most of the observed clusters. Thus, the UIT images provide a complete census of the hot UV-bright stars in the observed clusters. We have begun a program to obtain spectra of all the UV-bright stars found on the UIT images, in order to derive effective temperatures and gravities for the complete sample, for comparison with evolutionary tracks. Several of the UV-bright stars found on the UIT images, such as ROB 162 in NGC~6397, Barnard 29 in M~13, and vZ 1128 in M~3, were previously known and are well-studied. Other UIT stars are too close to the cluster cores for ground-based spectroscopy, and will require HST observations for further study. In this paper, we report on spectroscopy of those UIT UV-bright stars accessible for ground-based observations from the southern hemisphere. ", "conclusions": "The derived effective temperatures and gravities of the target stars are plotted in Figure 3, along with ZAHB and post-HB evolutionary tracks for [Fe/H] = $-$1.48 from Dorman et al. (\\cite{dorm93}), and post-AGB (0.565~\\Msolar) and post-early AGB (0.546~\\Msolar) tracks from Sch\\\"onberner (\\cite{scho83}). The stars NGC~6121-Y453, NGC~6723-III60, and NGC~6723-IV9 appear to fit the post-early AGB track, while the remaining four targets are consistent with post-EHB evolutionary tracks. In agreement with this scenario, the three post-early AGB stars have approximately solar helium abundances, while the post-EHB stars have subsolar helium abundances. The latter stars are expected to have subsolar helium abundances because they are direct descendants of EHB stars, which are known to show helium deficiencies (Moehler et al. \\cite{mohe97}), most likely due to diffusion processes. The post-early AGB stars, on the other hand, have evolved off the AGB, where the convective atmosphere is expected to eliminate any previous abundance depletions caused by diffusion. Curiously, no helium-rich (He/H $\\ge$ 1) sdO stars have yet been found in a globular cluster, although such stars dominate the field sdO population (Lemke et al. \\cite{lehe98}). As expected, the two clusters with a populous EHB (NGC~2808 and NGC~6752) have post-EHB stars but no post-AGB stars. The clusters NGC~6723 and M~4, on the other hand, do not have an EHB population, although they do have stars blueward of the RR Lyrae gap (which are potential progenitors of post-early AGB stars). The lack of true post-AGB stars may be understood from the different lifetimes: The lifetime of Sch\\\"onberner's post-early AGB track is about 10 times longer than his lowest mass post-AGB track. Thus, even if only a small fraction of stars follow post-early AGB tracks, those stars may be more numerous than true post-AGB stars. Due to their relatively long lifetime, post-early AGB stars are unlikely to be observed as central stars of planetary nebulae (CSPNe) since any nebulosity is probably dispersed before the central star is hot enough to ionize it. Additional detail on the individual stars is given below: \\subsection{NGC 2808} All three stars in NGC~2808 analysed in this paper are likely post-EHB stars. (Unfortunately, the best post-AGB candidate stars on the UIT image of NGC~2808 are too close to the cluster center to allow spectroscopy from the ground.) Although C2946 and C2947 could be separated in the long-slit optical spectra, they are too close together to estimate individual UV fluxes from the UIT image, and thus there is no UV luminosity determination in Table 2. Due to the well-populated EHB of NGC~2808 (Sosin et al., \\cite{sos97}), a large number of post-EHB stars are expected. From their three-colour WFPC2 photometry of NGC 2808, Sosin et al.\\ (\\cite{sos97}) find a larger distance modulus [(m-M)$_0$ = 15.25 -- 15.40] and lower reddening [\\Ebv\\ = 0.09 -- 0.16] than the values adopted here from Harris (\\cite{harris96}). The use of the distance and reddening of Sosin et al.\\ would yield masses about 20\\% larger, and luminosities about 0.05 dex larger than the values given in Table 2. \\subsection{NGC~6121 (M~4)} Y453 is possibly the hottest globular cluster star known so far. Other candidates are three central stars of planetary nebulae (IRAS 18333-2357 in M~22, Harrington \\& Paltoglou \\cite{hapa93}; JaFu1 in Pal~6 and JaFu2 in NGC~6441, Jacoby et al., \\cite{jamo97}), which however lack model atmosphere analyses of their stellar spectra. Such a high \\teff\\ is not unexpected, since according to the Sch\\\"onberner tracks, a post-AGB star will spend most of its lifetime at temperatures greater than 30,000~K. However, as pointed out by Renzini (\\cite{renz85}), the large bolometric corrections of such hot stars in the visible have biased the discovery of post-AGB stars in favour of cooler stars. In the \\logt --\\logg\\ plot, Y453 fits well on the 0.546~\\Msolar\\ post-early AGB track of Sch\\\"onberner (\\cite{scho83}). However, our derived luminosity (\\logl\\ = 2.6) is considerably lower than the Sch\\\"onberner track at that \\teff\\ and \\logg, and the derived mass of 0.16~\\Msolar\\ is astrophysically implausible. In order to obtain a mass of 0.55~\\Msolar, the value of \\logg\\ would need to be 5.68 instead of 5.15. This difference is too large to be accommodated by the spectral fitting, and still would not explain the discrepancy with the theoretically expected luminosity. Therefore, below we consider some other possible sources of error: \\begin{description} \\item [Line Blanketing:] The use of fully line-blanketed NLTE models for the analysis of Y453 might result in a somewhat lower temperature (0.05 dex) without any changes in surface gravity (Lanz et al. \\cite{lanz97}; Haas \\cite{haas97}). Such a lower temperature would increase the derived mass by about 15\\%. \\item [Differential Reddening:] Cudworth \\& Rees (\\cite{cud90}) find a gradient in the reddening that would increase the adopted reddening for Y453 (\\magpt{0}{35}) by about \\magpt{0}{015}. Lyons et al. (\\cite{lyon95}) report a patchiness in the reddening toward M~4 that is at least as significant as the gradient, and find a total range of 0.16 mag in \\Ebv . An increase by such a large amount would still lead to a mass of only 0.3~\\Msolar. Due to the non-standard reddening law toward M~4, the reddening correction for Y453 has a rather high uncertainty in any case. \\item [Distance:] The adopted distance (1.72 kpc) to M~4 is on the low side of the range of distance determinations but is supported by both a recent astrometric measurement (Rees \\cite{rees96}), and HST observations of the main-sequence (Richer et al.\\ \\cite{rich97}). Use of the distance given by Harris (\\cite{harris96}; 2200 pc) would give a mass of 0.26~\\Msolar. \\item [Photometry:] The only ground-based photometry of Y453 of which we are aware is the photographic photometry of Cudworth \\& Rees (\\cite{cud90}), who also derive a 99\\% probability of cluster membership from its proper motion. Y453 is among the faintest stars studied by Cudworth \\& Rees, so the photometric precision might be poorer than their quoted 0.025 mag (which corresponds to an error of 2\\% in M). \\end{description} \\subsection{NGC~6723} The V and B$-$V magnitudes in Table 1 for III-60 are from Menzies (\\cite{menz74}); we are not aware of any other photometry of this star. The tabulated photometry for IV-9 is from L.K. Fullton (1997, priv. comm.), who also gives U--B $= -0.84$. Photometry for IV-9 was also obtained by Menzies (\\cite{menz74}; V = 14.86, B$-$V $= -0.25$), and Martins \\& Fraquelli (\\cite{mafr87}; V=14.69, B$-$V $= -0.142$). III-60 and IV-9 fit well on the 0.546~\\Msolar\\ post-early AGB track (see above) of Sch\\\"onberner, and also have luminosities (\\logl $\\sim 3.0$) consistent with being post-early AGB stars. The spectrum of IV-9, however, was difficult to fit with any single model. As can be seen from Fig. 1 there is an absorption feature blueward of the \\ion{He}{I} line at 4713~\\AA , which matches the \\ion{He}{II} absorption line at 4686~\\AA\\ in wavelength. The \\ion{He}{II} line strength and the Balmer line profiles can be reproduced by a model with \\teff\\ = 30,000~K, \\logg\\ = 4.08, and \\loghe\\ = $-$0.89. However, this model is inconsistent with both the size of the Balmer jump and with the photometric indices (optical and UV) of this star. The photometric data indicate a temperature of 20000 -- 21000~K instead. Excluding the absorption feature at 4686~\\AA\\ from the fit results in a temperature of 20700~K and a \\logg\\ value of 3.34, in good agreement with the photometric temperature. The detection of metal lines (such as the \\ion{O}{II} absorption lines in BD+33$^\\circ$2642 discussed by Napiwotzki et al., \\cite{nahe93}) could help to decide between the two temperatures. Simulations with theoretical spectra, however, show that due to the low resolution of our data we cannot expect to see any metal lines. Any decision will therefore have to await better data. We keep the cooler temperature for all further analysis because of the good agreement with the photometric indices. \\subsection{NGC~6752} B2004 was one of only four post-EHB candidate stars present in the UIT color-magnitude diagram of NGC~6752 reported by Landsman et al. (\\cite{land96}), and the position of B2004 in the \\logt --\\logg\\ plot (Fig. 3) is consistent with post-EHB tracks. Landsman et al.\\ estimated \\teff\\ = 45000~K and \\logl\\ = 2.12 for B2004 on the basis of IUE spectrophotometry. However, the IUE photometry of B2004 had large uncertainties due to the presence of the nearby ($2.5''$ distant) blue HB star B1995, and the \\teff\\ (37000~K) and luminosity (\\logl\\ = 1.94) of B2004 derived here should be more accurate. Spectroscopic analyses of the other three post-EHB candidate stars (B852, B1754, and B4380) in NGC~6752 were presented by Moehler et al.\\ (\\cite{mohe97}). The four post-EHB stars in NGC~6752 occupy a fairly narrow range in temperature (4.5 $< \\log$ \\teff\\ $<$ 4.6) and luminosity ($1.94 < $ \\logl\\ $ < 2.12$), and are separated by a large luminosity gap (0.5~dex) from stars on the populous EHB. As discussed by Landsman et al.\\ (\\cite{land96}), these two characteristics are consistent with the non-canonical HB models of Sweigart (\\cite{sweig97}), which include helium mixing on the RGB. However, a more definitive test of EHB evolutionary tracks will require a larger sample of post-EHB stars." }, "9803/astro-ph9803125_arXiv.txt": { "abstract": "We describe an automated method for detecting clusters of galaxies in imaging and redshift galaxy surveys. The Adaptive Matched Filter (AMF) method utilizes galaxy positions, magnitudes, and---when available---photometric or spectroscopic redshifts to find clusters and determine their redshift and richness. The AMF can be applied to most types of galaxy surveys: from two-dimensional (2D) imaging surveys, to multi-band imaging surveys with photometric redshifts of any accuracy (2$\\half$D), to three-dimensional (3D) redshift surveys. The AMF can also be utilized in the selection of clusters in cosmological N-body simulations. The AMF identifies clusters by finding the peaks in a cluster likelihood map generated by convolving a galaxy survey with a filter based on a model of the cluster and field galaxy distributions. In tests on simulated 2D and 2$\\half$D data with a magnitude limit of $r' \\approx 23.5$, clusters are detected with an accuracy of $\\Delta z \\approx 0.02$ in redshift and $\\sim$10\\% in richness to $z \\lesssim 0.5$. Detecting clusters at higher redshifts is possible with deeper surveys. In this paper we present the theory behind the AMF and describe test results on synthetic galaxy catalogs. ", "introduction": "Clusters of galaxies---the most massive virialized systems known---provide powerful tools in the study of cosmology: from tracing the large-scale structure of the universe (\\cite{Bahcall88}, \\cite{Huchra90}, \\cite{Postman92}, \\cite{Dalton94}, \\cite{Peacock94} and references therein) to determining the amount of dark matter on Mpc scales (\\cite{Zwicky57}, \\cite{Tyson90}, \\cite{Kaiser93}, \\cite{Peebles93}, \\cite{Bahcall95}, \\cite{Carlberg96}) to studying the evolution of cluster abundance and its cosmological implications (\\cite{Evrard89}, \\cite{Peebles89}, \\cite{Henry92}, \\cite{Eke96}, \\cite{Bahcall97}, \\cite{Carlberg97}, \\cite{Oukbir97}). The above studies place some of the strongest constraints yet on cosmological parameters, including the mass-density parameter of the universe, the amplitude of mass fluctuations at a scale of 8~$\\hMpc$ and the baryon fraction. The availability of complete and accurate cluster catalogs needed for such studies is limited. One of the most used catalogs, the Abell catalog of rich clusters (\\cite{Abell58}, and its southern counterpart \\cite{Abell89}), has been extremely useful over the past four decades. This catalog, which contains $\\sim$4000 rich clusters to $z \\lesssim 0.2$ over the entire high latitude sky, with estimated redshifts and richnesses for all clusters, was constructed by visual selection from the Palomar Sky Survey plates, using well-defined selection criteria. The Zwicky cluster catalog (\\cite{Zwicky61}) was similarly constructed by visual inspection. The need for new, objective, and accurate large-area cluster catalogs to various depths is growing, following the important use of clusters in cosmology. Large area sky surveys using CCD imaging in one or several colors, as well as redshift surveys, are currently planned or underway, including, among others, the Sloan Digital Sky Survey (SDSS). Such surveys will provide the data needed for constructing accurate cluster catalogs that will be selected in an objective and automated manner. In order to identify clusters in the new galaxy surveys, a robust and automated cluster selection algorithm is needed. We propose such a method here. Cluster identification algorithms have typically been targeted at specific surveys, and new algorithms have been created as each survey is completed. \\cite{Abell58} was the first to develop a well-defined method for cluster selection, even though the identification was carried out by visual inspection (see, e.g., \\cite{McGill90} for a analysis of this method). Other algorithms have been created for the APM survey (\\cite{Dalton94}, \\cite{Dalton97}; see \\cite{Schuecker98} for a variant of this method), the Edinburgh-Durham survey (ED; \\cite{Lumsden92}), and the Palomar Distant Cluster Survey (PDCS; \\cite{Postman96}; see also \\cite{Kawasaki97} for a variant of this method; and \\cite{Kleyna97} for an application this method to finding dwarf spheroidals). All the above methods were designed for and applied to two-dimensional imaging surveys. In this paper we present a well defined, quantitative method, based on a matched filter technique that expands on some of the previous methods and provides a general algorithm that can be used to identify clusters in any type of survey. It can be applied to 2D imaging surveys, 2$\\half$D surveys (multi-band imaging with photometric redshift estimates of any accuracy), 3D redshift surveys, as well as combinations of the above (i.e. some galaxies with photometric redshifts and some with spectral redshifts). In addition, this Adaptive Matched Filter (AMF) method can be applied to identify clusters in cosmological simulations. The AMF identifies clusters by finding the peaks in a cluster likelihood map generated by convolving a galaxy survey (2D, 2$\\half$D or 3D) with a filter which models the cluster and field galaxy distribution. The peaks in the likelihood map correspond to locations where the match between the survey and the filter is maximized. In addition, the location and value of each peak also gives the best fit redshift and richness for each cluster. The filter is composed of several sub-filters that select different components of the survey: a surface density profile acting on the position data, a luminosity profile acting on the apparent magnitudes, and, in the 2$\\half$D and 3D cases, a redshift cut acting on the estimated redshifts. The AMF is adaptive in three ways. First, the AMF adapts to the errors in the observed redshifts (from no redshift information (2D), to approximate (2$\\half$D) or measured redshifts (3D)). Second, the AMF uses the location of the galaxies as a ``naturally'' adaptive grid to ensure sufficient spatial resolution at even the highest redshifts. Third, the AMF uses a two step approach that first applies a coarse filter to find the clusters and then a fine filter to provide more precise estimates of the redshift and richness of each cluster. We describe the theory of the AMF in \\S2 and its implementation in \\S3. We generate a synthetic galaxy catalog to test the AMF in \\S4 and present the results in \\S5. We summarize our conclusions in \\S6. ", "conclusions": "We have presented the Adaptive Matched Filter method for the automatic selection of clusters of galaxies in a wide variety of galaxy catalogs. The AMF can find clusters in most types of galaxy surveys: from two-dimensional (2D) imaging surveys, to multi-band imaging surveys with photometric redshifts of any accuracy (2$\\half$D), to three-dimensional (3D) redshift surveys. The method can also be utilized in the selection of clusters in cosmological N-body simulations. The AMF is based on matching the galaxy catalog with a cluster filter that models the overall galaxy distribution. The model describes the surface density, apparent magnitude, and redshift of cluster and field galaxies. Convolving the data with the filter produces a cluster probability map whose peaks correspond to the location of the clusters. The probability peaks also yield the best fit redshift and richness of each cluster. The heart of the AMF is the apparent overdensity $\\delta_i$ which is evaluated at each galaxy position and has a higher value for galaxies in clusters than galaxies in the field. The apparent overdensity distills the entire description of the galaxy catalog into a single function. Two likelihood functions are derived, $\\sL_\\coarse$ and $\\sL_\\fine$, using different underlying model assumptions. The theoretical framework of the AMF allows estimated redshifts to be included via a simple redshift filter, which effectively limits the sums in $\\sL_\\coarse$ and $\\sL_\\fine$ to those galaxies within a window around $z_c$. The maxima in the likelihood functions are used to identify cluster positions as well as their redshifts and richnesses. The AMF is adaptive in three ways. First, it adapts to the errors in the estimated redshifts. Second, it uses the locations of the galaxies as ``naturally'' adaptive grid to ensure sufficient resolution at even the highest redshifts. Third, it uses a two step approach that applies a coarse filter to initially find the clusters and a fine filter to more precisely estimate the redshift and richness of each cluster. We tested the AMF on a set of simulated clusters with different richnesses and redshifts---ranging from groups to rich clusters at redshifts 0.1 to 0.5; the clusters were placed in a simulated field of randomly distributed galaxies as well as in a non-random distribution produced by N-body cosmological simulations. We find that the AMF detects clusters with an accuracy of $\\Delta z \\sim$0.02 in redshift and $\\sim$10\\% in richness to $z \\lesssim 0.5$ (for a simulated galaxy survey to $r' \\approx 23.5$). In addition, robustness tests provide a strong indication that the AMF will perform well on observational data sets. Detecting clusters at even higher redshifts will be possible in deeper surveys." }, "9803/astro-ph9803038_arXiv.txt": { "abstract": "We present a new method for reconstructing two-dimensional mass maps of galaxy clusters from the image distortion of background galaxies. In contrast to most previous approaches, which directly convert locally averaged image ellipticities to mass maps (direct methods), our entropy-regularized maximum-likelihood method is an inverse approach. Albeit somewhat more expensive computationally, our method allows high spatial resolution in those parts of the cluster where the lensing signal is strong enough. Furthermore, it allows to straightforwardly incorporate additional constraints, such as magnification information or strong-lensing features. Using synthetic data, we compare our new approach to direct methods and find indeed a substantial improvement especially in the reconstruction of mass peaks. The main differences to previously published inverse methods are discussed. ", "introduction": "The reconstruction of projected cluster mass maps from the observable image distortion of faint background galaxies due to the tidal gravitational field is a new and powerful technique. Pioneered by Kaiser \\& Squires (1993), this method has since been modified and generalized to account for (a) strong tidal fields in cluster centers (Schneider \\& Seitz 1995; Seitz \\& Schneider 1995; Kaiser 1995); (b) finite and -- in some cases, e.g.~WFPC2 images -- very small data fields (Schneider 1995; Kaiser et al.\\ 1995; Bartelmann 1995; Seitz \\& Schneider 1996, 1998; Lombardi \\& Bertin 1998); and (c) the broad redshift distribution of background galaxies (Seitz \\& Schneider 1997). All of these are direct methods in the sense that a local estimate of the tidal field is derived from observed galaxy ellipticities, which is then inserted into an inversion equation to obtain an estimate of the surface mass density of the cluster. Whereas these direct methods are computationally fast, can be treated as black-box routines, need only the observed ellipticities and a smoothing length $\\theta_0$ as input data, and yield fair estimates of the surface mass density, their application has several drawbacks: \\begin{itemize} \\item The data must be smoothed, and the smoothing scale is typically a free input parameter specified prior to the mass reconstruction. There are no objective criteria on how to set the smoothing scale, although some ad-hoc prescriptions for adapting it to the strength of the lensing signal have been given (Seitz et al.\\ 1996). In general, smoothing leads to an underestimate of the surface mass density in cluster centers or sub-condensations. \\item The quality of the reconstruction is hard to quantify. \\item Constraints on the mass distribution from additional observables (such as multiple images or giant arcs) cannot simultaneously be included. In particular, magnification information contained in the number density of background sources (Broadhurst et al.\\ 1995; Fort et al.\\ 1997) or in the image sizes at fixed surface brightness (Bartelmann \\& Narayan 1995), cannot be incorporated locally but only globally to break the mass-sheet degeneracy (Gorenstein et al.\\ 1988; Schneider \\& Seitz 1995). \\end{itemize} To overcome these drawbacks, a different class of methods should be used. Bartelmann et al.\\ (1996, hereafter BNSS) developed a maximum-likelihood (ML) technique in which the values of the deflection potential at grid points are considered as free parameters. After averaging image ellipticities and sizes over grid cells, local estimates of shear and magnification are obtained. The deflection potential at the grid points is then determined such as to optimally reproduce the observed shear and magnification estimates. Magnification information can be included this way. The smoothing scale in this method is given by the size of the grid cells, and can be chosen such that the overall $\\chi^2$ of the fit is of order unity per degree of freedom. Squires \\& Kaiser (1996; hereafter SK) suggested several inverse methods. Their {\\em maximum probability method\\/} parameterizes the mass distribution of the cluster by a set of Fourier modes. If the number of degrees of freedom (here the number of Fourier modes) is large, the mass model tends to over-fit the data. This has to be avoided by regularizing the model, for which purpose SK impose a condition on the power spectrum of the Fourier modes. SK's {\\em maximum-likelihood method\\/} specifies the surface mass density on a grid and uses the Tikhonov-Miller regularization (Press et al.\\ 1992, Sect.\\ 18.5). The smoothness of the mass reconstructions can be changed by varying the regularization parameter, which is chosen such as to give an overall $\\chi^2\\approx1$ per degree of freedom. Bridle et al.\\ (1998) have recently proposed an entropy-regularized ML method in which the cluster mass map is parameterized by the surface mass density at grid points. This method allows to restrict the possible mass maps to such with non-negative surface mass density. This paper describes another variant of the ML method (Seitz 1997, Ph.D.\\ thesis). The major differences to the previously mentioned inverse methods are the following: \\begin{itemize} \\item The observational data (e.g.~the image ellipticities) are not smoothed, but each individual ellipticity of a background galaxy is used in the likelihood function. Whereas this modification complicates the implementation of the method, it allows larger spatial resolution for a given number of grid points, which is useful since the latter determines the computing time. \\item The number of grid points can be much larger than in BNSS, and the likelihood function is regularized. This produces mass reconstructions of variable smoothness: Mass maps are smooth where the data do not demand structure, but show sharp peaks where required by the data. The resulting spatially varying smoothing scale is a very desirable feature. Fourier methods, such as SK's maximum probability method, have a spatially constant smoothing scale which is determined by the highest-order Fourier components. They always need to compromise between providing sufficient resolution near mass peaks and avoiding over-fitting of the data in the outer parts of a cluster. \\item Following BNSS, we use the deflection potential to describe a cluster. This is an essential difference to Bridle et al.\\ (1998) who used the surface mass density at grid points. As we shall discuss below, working with the deflection potential has substantial fundamental and practical advantages. \\end{itemize} We describe our method in Sect.\\ 2, with details given in the Appendix. We then apply the method to synthetic data sets in Sects.\\ 3 \\& 4 to demonstrate its accuracy. In particular, we compare the performance of our ML method to that of direct methods. The results are then discussed in Sect.\\ 4, and conclusions are given in Sect.\\ 5, where we also discuss further generalizations of the method for, e.g., including constraints from strong lensing features. ", "conclusions": "We presented a new method for reconstructing projected mass distributions of galaxy clusters. The method uses image distortions of background galaxies and their size as a function of surface brightness. Our entropy-regularized ML method (Seitz 1997) is a further development of previously published inverse methods for the mass reconstruction. In particular, we describe the lens by its deflection potential $\\psi$ as suggested by BNSS. This is of major importance, for two reasons. First, if the surface mass density $\\kappa$ on a finite field $\\U$ is used to describe the lens, the shear on $\\U$ is incompletely specified by the model because the mass distribution outside $\\U$ can contribute to the shear. Second, the shear at the position of any galaxy depends only locally on $\\psi$, which allows a much faster minimization algorithm for a given number of grid points. We regularize the method by an entropy term as suggested by Bridle et al.\\ (1998), but additionally adapt the prior to the current model of the mass distribution. This `moving prior' (Lucy 1994) allows a considerably higher resolution of mass peaks. The spatial resolution of the entropy-regularized ML method adapts itself to the strength of the lensing signal, producing mass distributions which are as smooth as possible, and as structured as the data require. In that respect, our method differs from that of BNSS and SK. We showed that the ML method is superior to the noise-filter method (Seitz \\& Schneider 1996) which was the most accurate of the presently known direct inversion methods (Seitz \\& Schneider 1996, 1998; SK; Lombardi \\& Bertin 1998). Obviously, the method described here is not restricted to rectangular data fields, but can easily be adapted to any geometry of $\\U$ by covering $\\U$ with quadratic grid cells, and adding a boundary of grid points for $\\psi$ -- the rest is only a matter of labeling. Furthermore, we note that observational errors can be incorporated into the likelihood function. For example, if the measurement error of the ellipticity $\\chi$ is $\\sigma_{\\rm obs}$, one can replace $\\sigma_\\chi^2$ in (\\ref{eq:2.13} and \\ref{eq:2.14}) by $\\sigma_\\chi^2+\\sigma_{\\rm obs}^2$. In contrast to the direct inversion methods, all of which are variants and generalizations of the original Kaiser \\& Squires (1993) method, the inverse methods allow to include additional information on top of the shear measured through image ellipticities. We demonstrated this here by adding magnification information derived from image sizes at given surface brightness, as discussed by Bartelmann \\& Narayan (1995). However, we could equally well use the change of number counts due to magnification bias (Broadhurst et al.\\ 1995) as an additional constraint. In that case, if the number counts of a certain (e.g.~color-selected) galaxy population have a cumulative slope of $-\\beta$, the expected number density of background galaxies at a position $\\vc\\theta$ is $n(\\vc\\theta)=n_0\\abs{\\mu(\\vc\\theta)}^{\\beta-1}$, where $n_0$ are the counts at the same flux limit in the absence of lensing. Assuming that galaxies are intrinsically randomly distributed, the probability of having $N$ galaxies within $\\U$ is a Poisson distribution $P_N(\\ave{N})$ with \\begin{equation} \\ave{N} = n_0\\int_\\U\\,\\d^2\\theta\\, \\abs{\\mu(\\vc\\theta)}^{\\beta-1}\\;. \\label{eq:5.1} \\end{equation} Consequently, the likelihood function could be augmented by a factor \\begin{equation} \\L_\\mu = P_N(\\ave{N})\\prod_{k=1}^N\\, \\abs{\\mu(\\vc\\theta_k)}^{\\beta-1}\\;. \\label{eq:5.2} \\end{equation} If galaxy clustering is important, the likelihood $\\L_\\mu$ cannot be written as a simple product over individual galaxies, but the joint probability distributions must be taken into account. The contribution of clustering effects to the likelihood function is somewhat uncertain, because an approximate expression has to be used due to lack of knowledge on the $N$-point correlation functions (see Broadhurst et al.\\ 1995 for further discussion). Perhaps the most promising generalization of our method is the inclusion of strong lensing constraints. Since giant arcs and multiple images of background galaxies provide (nearly) exact constraints on the lens mass distribution, it is highly desirable to include them into a mass reconstruction. The obvious way to do this would be to augment the function $E$ by a term which measures the degree to which multiple images of the same source are mapped back to the same position in the source plane. In addition, the surface brightness profile of multiple images of extended sources can be incorporated, e.g.~in a similar manner as the spatially resolved multiple arc in the cluster Cl0024+16 (Colley at al.\\ 1996). In some of the observed clusters, the lensing effects of individual galaxies are visible, in particular through deformations of giant arcs. Some of the most prominent examples are the triple arc in 0024+16 (Kassiola et al.\\ 1992), the multiple arc systems in A~2218 (Kneib et al.\\ 1996), and the distortion of the curvature in the arc of the galaxy cB58 in MS1512+36 (Seitz et al.\\ 1998). But even weaker lensing effects of individual (cluster) galaxies can be detected using a combination of cluster mass reconstruction and galaxy-galaxy lensing techniques. By adding two free parameters to the lens model, such as the mass-to-light ratio of cluster galaxies and their characteristic spatial extent, the size of halos of cluster galaxies can be investigated (Natarayan et al.\\ 1997; Geiger \\& Schneider 1998)." }, "9803/astro-ph9803206.txt": { "abstract": "Differences between observed and theoretical eigenfrequencies of the Sun have characteristics which identify them as arising predominantly from properties of the oscillations in the vicinity of the solar surface: in the super-adiabatic, convective boundary layer and above. These frequency differences may therefore provide useful information about the structure of these regions, precisely where the theory of solar structure is most uncertain. In the present work we use numerical simulations of the outer part of the Sun to quantify the influence of turbulent convection on solar oscillation frequencies. Separating the influence into effects on the mean model and effects on the physics of the modes, we find that the main model effects are due to the turbulent pressure that provides additional support against gravity, and thermal differences between average 3-D models and 1-D models. Surfaces of constant pressure in the visible photosphere are elevated by about 150 km, relative to a standard envelope model. As a result, the turning points of high-frequency modes are raised, while those of the low-frequency modes remain essentially unaffected. The corresponding gradual lowering of the mode frequencies accounts for most of the frequency difference between observations and standard solar models. Additional effects are expected to come primarily from changes in the physics of the modes, in particular from the modulation of the turbulent pressure by the oscillations. ", "introduction": "In standard solar models, the stratification of the convection zone is determined by mixing-length theory (MLT), thereby reducing the entire complexity of the turbulent hydrodynamics to a one-parameter family of models. MLT solar models suffer from several basic inconsistencies. For example, they predict that convective velocities, of several km$\\,$s$^{-1}$, disappear abruptly in a few tens of km, immediately below the solar surface. %at the edge of the convective zone. %Velocities are predicted to disappear In contrast, observations show convective cells with those same characteristic velocities, and with horizontal sizes of 1000 -- 5000 km, whose velocity fields obviously cannot vanish so abruptly. Indirect evidence from spectral line broadening indeed shows that the photosphere is pervaded by a velocity field with rms Mach numbers of the order of 0.3, and yet, in standard solar models these layers are assumed to be entirely quiescent. Clearly, such a discrepancy between the theoretical description and the observations should be regarded as a warning not to take the quantitative predictions of the theory too seriously. Helioseismology provides quantitative diagnostics that pertain precisely to these critical surface layers, since this is where the upper turning points of the majority of modes are located. Thus, analysis of the observed properties of these modes may help clarify the consequences of the inconsistencies inherent in MLT and its more recent siblings (\\cite{CM91,CM92,Canuto+Goldman+Mazzitelli96}). Indeed, as we discuss in more detail below, adiabatic oscillations of MLT models show significant systematic discrepancies when compared with measured solar frequencies. It is natural, therefore, to seek to use helioseismology applied to these differences to improve the theoretical description. However, this procedure is undermined by our present uncertainty about the physics of the oscillations near the top of the convection zone where they are likely to be strongly coupled to both the convective and radiative fields. In the language of \\citetext{Balmforth92b}, the {\\emph extrinsic} (or {\\emph model}\\/) error in the mode frequencies (due to inaccurate modeling of the mean solar structure) cannot be accurately estimated while the {\\emph intrinsic} (or {\\emph modal}\\/) errors (due to uncertain mode physics) are largely unknown. %\\note [I have replaced extrinsic and intrinsic by model and %modal throughout; however, we may need another sentence here %to hammer the notation down. !jcd] One approach to resolving this problem has been the time-dependent non-local non-adiabatic mixing-length theory of \\citetext{Gough77} and Balmforth (1992abc) in which the \\nocite{Balmforth92a,Balmforth92b,Balmforth92c} coupling of the oscillations to both the convection and the radiation is included within the framework of mixing-length theory. Another approach to the problem has been proposed by \\citetext{Zhugzhda+Stix94} who have developed a model of the modal effect on mode frequencies due to advection of the oscillations by spatially varying radial flows. This approach has not yet been developed to the stage where it can be usefully applied to realistic solar models with stratification and turbulent pressure. Finally, \\citetext{Rudiger+97} have used turbulence closure assumptions to parameterize the propagation of acoustic disturbances through a convecting medium. In the present work, we use an alternative technique for estimating the model effects, based on the results obtained from numerical simulations of turbulent convection in a radiating fluid. We show that p modes can be calculated from a mean model with hydrostatic and thermodynamic stratification obtained by appropriately weighted averages of the simulation results. We proceed by making simplifying, and certainly very naive, assumptions about the modal effects, postponing their detailed study to subsequent papers. We begin (Section 2) with a brief discussion of the helioseismic data and the discrepancy between measured frequencies and those calculated from MLT models. We then discuss (Section 3) the averaging techniques needed to analyze the radial oscillations of a convecting layer, describe the model computations (Section 4), and investigate the resulting oscillation frequencies (Section 5). Finally (Section 6), we discuss the relevance and limitations of the results, and indicate future plans. ", "conclusions": "In the classical theory of solar structure, a one-parameter family of models (MLT) is calibrated against the known radius of the Sun. It is well known that MLT is based on a number of fundamentally inapplicable and inconsistent assumptions, and so a large number of alternative models of stellar convection have been proposed. Non-local mixing-length theory (\\cite{Gough77}; Balmforth 1992abc\\nocite{Balmforth92a,Balmforth92b,Balmforth92c}) attempts to improve on standard MLT by incorporating into it the effect of the finite size of turbulent eddies. \\citetext{Forestini+91} have produced an MLT-type model incorporating a measure of asymmetry between upflows and downflows as found in the simulations. The models of \\citeauthor{Lydon+92} (\\citeyear{Lydon+93a}, \\citeyear{Lydon+93b}) are essentially attempts to parameterize a wide range of convection simulations using a formalism similar to MLT but incorporating also the contribution of the kinetic energy flux to the flux-balance equation. \\citetext{CM91} have taken an approach based on modern theories of turbulence, and have attempted to produce a parameterized expression based on such theories. Finally, Canuto (\\citeyear{Canuto92,Canuto93}) has produced an ambitious model of convection based on a Reynolds' stress formalism. These more elaborate models of convection are potentially very useful, if it can be shown that they capture essential aspects of the full 3-D convection in terms of much simpler equations. In particular, non-local and time-dependent models of convection such as the ones by \\citetext{Gough77}, Buchler (1993), and Houdek (1997) would be very \\nocite{Houdek97,Buchler93} useful, for example in the modeling of pulsating stars, since full 3-D simulations are too expensive to be used in that context. The most obvious way to proceed is by use of numerical simulations such as the one used here or those of \\citetext{Kim+95}, treating the simulations as data against which the models are to be tested and calibrated. A second approach to testing the simplified models, and one which has been more widely adopted so far, is the helioseismic approach in which the frequencies of oscillations of a solar model constructed with a given theory are compared with the observed frequencies. Such a procedure, however, must deal with the difficulty that the complete structure of the surface layers is certainly underdetermined by the (low- and intermediate-degree) oscillation frequencies since, as \\citetext{JCD+Thompson97} have shown, the near-surface contribution to the frequencies depends only on $\\upsilon$ and not separately on, e.g., $p_0$ and $\\Gamma_1$. Thus improved agreement with measured mode frequencies cannot, by itself, be taken as evidence that a given model of convection is a better description of reality. In the present work we have attempted to improve on this approach by analyzing the problem in such a way as to obtain a physically justifiable description of the oscillations. Moreover, by using a numerical simulation of convection we give ourselves no free parameters with which to calibrate our model. Given this approach, the fact that we are able to construct a complete solar envelope model with essentially the correct convection-zone depth is, in itself, a considerable achievement for the simulation. We have here investigated model effects on the mode frequencies, primarily those due to changes in pressure support of the atmosphere and 3-D radiation transfer. The effects we have found are robust; there is no question that the averaging of 3-D fluctuations results in differences of this sense and order of magnitude; the turbulent pressure elevation is constrained by the observed photospheric velocity field, and the mean thermal difference is an inevitable consequence of the temperature dependence of the opacity. One might attempt to estimate the elevation effect from turbulent pressure using a local convection model such as MLT or the model of Canuto \\& Mazzitelli (1991, 1992). However, as discussed by \\citetext{Antia+Basu97score}, such models cannot accurately account for the effect of turbulent pressure because they result in an artificially abrupt upper boundary of the convection zone and therefore a serious overestimate of the turbulent-pressure gradient there. In addition, as discussed in section \\ref{pturb.sec}, a 1-D model with the correct pressure stratification unavoidably has a surface radiation flux that corresponds to an incorrect effective temperature. Thus, there are inherent limitations in simplified 1-D models of convection. The principal remaining uncertainty in determining the oscillation frequencies from 3-D models lies in the uncertain mode physics. In particular, non-adiabatic effects, and the response of the turbulent pressure to the compression and rarefaction in the oscillations needs to be understood. % We have attempted to quantify this uncertainty by calculating two models: the RGM in which the turbulent pressure is assumed to be unaffected by the oscillations and the GGM in which it is assumed to respond in exactly the same way as does the gas pressure. The frequency discrepancies for the RGM are almost exactly twice those for the GGM. The GGM produces frequencies that are closer to those observed, but this should of course not be taken as evidence that the turbulent pressure responds in exactly the same way as the gas pressure. The actual depth- and time-dependent mode response of the turbulent pressure produces, together with the response of the gas pressure, a complex-valued, and frequency-dependent $\\Gamma_1(z,\\omega)$. % At any particular frequency, the real part of $\\Gamma_1$ may be expected to have a different depth dependence than that assumed in both the RGM and GGM models; therein lies an essential part of the modal effects, and thus a potential explanation for part of the remaining differences between the observed and calculated oscillation frequencies. Preliminary investigations of the relation between $\\delta \\ln \\bra \\rho \\ket$ and $\\delta \\ln \\bra P \\ket$ in numerical 3-D simulations overlaid with initial radial eigenmodes show that non-adiabatic effects are indeed significant. The effective gamma appears to be closer to unity than to $5/3$ in the optically thin parts of the photosphere, while in the very surface layers the effective $\\Gamma_1$ can become quite large ($\\sim 8$), because of a localized reduction of $\\delta\\rho$. A more quantitative analysis of the non-adiabatic effects will require much more work, though, and will appear in a subsequent paper. Additional differences (in particular the ones reflected in the discrepancy of the f-mode frequencies) are expected to come from true 3-D effects; differences between mode behavior in a homogeneous and inhomogeneous medium. Again, a quantitative investigation of such effects requires elaborate, differential comparisons between non-radial mode behavior in 1-D and 3-D models, and is beyond the scope of the present paper. However, a pre-requisite for studying mode physics effects is a sufficient accuracy of the mean model; only if one includes the model effects with sufficient accuracy does it make sense to use remaining discrepancies to diagnose modal effects. How accurate, then, is the pressure stratification obtained from the present numerical simulations? The tests at various numerical resolutions show that the location, shape, and width of the peak of the turbulent pressure (relative to the total pressure) are quite insensitive to the numerical resolution, while the magnitude of the turbulent pressure increases slightly with increasing numerical resolution (\\Fig{F8}). The magnitude scaling of the turbulent pressure is, however, tightly constrained by spectral line-broadening observations (\\Fig{F9}). The existence of turbulent pressure support of about the magnitude found here thus cannot be doubted. Moreover, part of the elevation effect is due to the thermal difference between 1-D and average 3-D models. Any calculation of mode frequencies that does not include a turbulent and thermal pressure elevation of the upper turning points of the modes is thus neglecting a significant effect. If parameter fitting for such a calculation leads to near agreement with the observations it merely illustrates that it is quite possible to obtain ``the right result for the wrong reason''. Finally we must emphasize that while our understanding of convective effects on radial oscillations may seem rudimentary, it is in fact highly sophisticated by comparison with our understanding of their influence on nonradial modes. Indeed if we consider the most nonradial mode of all, the f mode, in which radial and nonradial motions are of equal magnitude, we note that no explanation which seeks to replace convection modeling with a hydrostatic solar model can ever explain the measured frequency residuals because f-mode frequencies are largely insensitive to hydrostatic structure. Thus, both the f-mode frequency discrepancies and the remaining discrepancies in \\Fig{mobsdiff} are likely to be caused by modal effects, rather than by the stratification effects that we have uncovered in the present paper. Evidently the behaviour of both nonradial and radial modes needs more study. In order to address the modal effects on nonradial modes it will be necessary to invoke a more elaborate technique than the simple planar averages we have used, a result already evident from the simplified model calculations of \\citetext{Zhugzhda+Stix94}. On the one hand this emphasizes the gulf which still exists between theory and measurement in mode physics but, on a more positive note, it suggests that the remaining frequency discrepancies may have great power as diagnostic probes of the structure of turbulent convection." }, "9803/nucl-th9803026_arXiv.txt": { "abstract": "The properties of hot matter are studied in the frame of the relativistic Brueckner-Hartree-Fock theory. The equations are solved self-consistently in the full Dirac space. For the interaction we used the potentials given by Brockmann and Machleidt. The obtained critical temperatures are smaller than in most of the nonrelativistic investigations. We also calculated the thermodynamic properties of hot matter in the relativistic Hartree--Fock approximation, where the force parameters were adjusted to the outcome of the relativistic Brueckner--Hartree--Fock calculations at zero temperature. Here, one obtains higher critical temperatures, which are comparable with other relativistic calculations in the Hartree scheme. ", "introduction": "The properties of hot and dense nuclear matter play an essential role in the understanding of high-energy heavy-ion collisions, supernova explosions and proto-neutron stars. For that reason the problem of hot nuclear matter has been studied over the last decades in several investigations, which however were predominantly performed within the nonrelativistic scheme \\cite{1a}, using either effective density dependent interactions \\cite{1a,1,2,23} or the Brueckner approach \\cite{1a,3,4,5,5a}. In the relativistic approach investigations of the equations of state for $T\\neq 0$ are relatively scarce. The majority of such calculations were performed in the relativistic Hartree approximation (RH), where the extension to finite temperatures is straightforward. Details of this scheme are given, for instance, in Refs.\\,\\cite{23}, \\cite{6}-\\cite{11}. More complicated are the relativistic Hartree--Fock-- \\cite{13,11} and the Brueckner--Hartree--Fock approximation \\cite{12}, and the application to finite nuclei \\cite{14}. In this contribution we will concentrate on the relativistic Brueckner--Hartree--Fock treatment (RBHF) of symmetric and asymmetric nuclear matter generalizing the formalism as described in Refs.\\,\\cite{15,15a} to $T\\neq 0$. The RBHF--approach seems to be of special interest, since it is known for $T = 0$ that the resulting EOSs are much stiffer than their nonrelativistic counterparts \\cite{15b}. To our knowledge such an investigation has been only performed so far by the Groningen group for symmetric nuclear matter \\cite{12}. As described in more details in Refs.\\,\\cite{12,15}, their method uses a nonunique ansatz for the T--matrix in terms of five Fermi invariants, which can lead to ambiguous results for the self-energies (see, e.g., Refs.\\,\\cite{17,18}). In order to avoid this problem we solve, according to the original scheme of the Brooklyn group \\cite{19}, the RBHF--approximation in the full Dirac space, which is more tedious (for a more detailed discussion, see Ref.\\,\\cite{15}). Since the formal structure of the problem is the same as for $T=0$, where one has to solve three coupled equations, namely the Dyson equation for the one-body Green's function $G$, the (reduced) Bethe--Salpeter equation for the effective scattering matrix $T$ in matter and the equation for the self--energy $\\Sigma$, we will not repeat here the equations. As in Refs.\\,\\cite{12,15,15a} we will restrict ourselves to the incorporation of intermediate positive--energy nucleon states only, where now the Fermi step functions are replaced by the Fermi distribution functions $n_{\\vec p}\\,(T)$ at finite temperature $T$ (for details, see Ref.\\,\\cite{21}). The Green's function obeys for $T\\neq 0$ the spectral representation \\cite{11,20} \\begin{equation} \\label{I.1} G(p) = \\int d\\omega\\,A(\\vec{p},\\omega) \\left\\{\\frac{f(\\omega)}{p_\\rho -\\omega - i\\eta} + \\frac{f(-\\omega)}{p_\\rho -\\omega+i\\eta}\\right\\} ~, \\end{equation} with \\begin{equation} \\label{I.2} f(\\omega) = (e^{\\beta\\omega}+1)^{-1} ~ {\\stackrel{T=0}{_{\\mbox{$\\longrightarrow$}}}} ~ \\Theta(-\\omega) ~. \\end{equation} The formal structure of the spectral function $A(\\vec{p},\\omega)$ \\cite{15} remains unaltered to the case for $T=0$. A further difference in comparison with Ref.\\,\\cite{12} is that we take the momentum dependence of the self--energies into account. Since the pole of the quasi--particle propagators occurs for $T\\neq 0$ in the integration domain of the intermediate states, one obtains, in principle, complex effective scattering matrix elements and self--energies. It was checked in Ref.\\,\\cite{12} that the imaginary part of $\\Sigma$ turned out to be small. Therefore we neglect also Im~$\\Sigma$ in the calculations. For the one--boson--exchange interaction we used the modern potentials constructed by Brockmann and Machleidt \\cite{22}. We select for the presentation the so-called potential $B$, which gives the best results for the EOS ($E/A = -15.73$~MeV; $\\rho_0 = 0.172$~fm$^{-3}; K = 249$~MeV; $J = 32.8$~MeV) at zero temperature in RBHF--calculations (see Refs.\\,\\cite{15,15a}, for the potential $A$ the outcome is similar, see Ref.\\,\\cite{21}). For the sake of comparison we also treated the RHF--approximation, where we adjusted the force parameters to the outcome for the EOS for symmetric and asymmetric nuclear matter at $T=0$ \\cite{15a,21}. The RHF--approximation has in comparison with the RH--approximation the advantage to resemble in its formal structure more to the RBHF--approximation with the benefit of a much simpler numerical treatment than in the RBHF case. For finite temperatures one needs for the determination of the pressure the free energy per baryon, defined as \\begin{equation} \\label{I.3} f = u - T s ~, \\end{equation} with the entropy per baryon: \\begin{equation} \\label{I.4} s = - \\frac{2}{\\rho h^3} \\sum_\\tau \\int d^3p\\,[n(\\vec{p})\\; {\\rm ln} \\,n(\\vec{p}) + \\left(1 - n(\\vec{p}\\right)\\; {\\rm ln}\\,\\left(1 - n(\\vec{p}\\right)]~. \\end{equation} The pressure is given by: \\begin{equation} \\label{I.5} P = \\rho \\sum_\\tau \\rho_\\tau \\left(\\frac{\\partial f} {\\partial\\rho_\\tau}\\right)_{T,\\rho_{-\\tau}} ~. \\end{equation} ", "conclusions": "In conclusion, we have performed a calculation of hot symmetric and asymmetric nuclear matter within the relativistic Brueckner--Hartree--Fock scheme using modern OBE--interactions constructed by Brockmann and Machleidt. It turned out that the critical temperatures are smaller than it is the case for the majority of nonrelativistic treatments. We have additionally treated the relativistic Hartree--Fock approximation at $T\\neq 0$, where the Lagrangian parameters were adjusted to the outcome of the RBHF--treatment for $T=0$. Here the critical temperature is in the range of other relativistic treatments performed in the Hartree scheme." }, "9803/astro-ph9803042_arXiv.txt": { "abstract": "We report on the discovery of a new kind of thermal pulse in intermediate mass AGB stars. Deep dredge-up during normal thermal pulses on the AGB leads to the formation of a long, unburnt tail to the helium profile. Eventually the tail ignites under partially degenerate conditions producing a strong shell flash with very deep subsequent dredge-up. The carbon content of the intershell convective region (X$_{C}$ $\\sim$ 0.6) is substantially higher than in a normal thermal pulse (X$_{C}$ $\\sim$ 0.25) and about 4 times more carbon is dredged-up than in a normal pulse. ", "introduction": "It is now just over 30 years since thermal pulses were discovered in AGB stars by \\cite{Sch65} and over 20 years since third dredge-up was discovered by \\cite{Ibe75}, and we are still learning about the consequences of these events for stellar evolution and nucleosynthesis. Although much is known there are still many uncertainties, especially concerning third dredge-up (\\cite{Fro96}) and mass-loss. We are presently studying the effects of hot bottom burning (HBB) on intermediate mass AGB stars during their thermally pulsing evolution. During these calculations we found a new kind of thermal pulse which we report on in this paper. ", "conclusions": "We believe that the development of degenerate thermal pulses is inevitable provided two criteria are met:~1)~dredge-up is very deep, providing cooling of the helium shell soon after ignition;~and~2)~the star lives long enough on the AGB for the helium shell to reach (partially) degenerate conditions. Whether these conditions are realised in real stars remains an open question. The two aspects of AGB evolution which are the most uncertain are the extent of dredge-up and the rate of mass-loss (hence AGB lifetime), precisely the two phenomena which govern the occurrence or not of degenerate thermal pulses. It seems prudent to look for some observational consequence of degenerate thermal pulses. The nucleosynthetic consequences are currently being investigated, and will be reported elsewhere. \\clearpage" }, "9803/astro-ph9803274_arXiv.txt": { "abstract": "Deep $J, H,$ and $K$ images are used to probe the evolved stellar contents in the central regions of the Sculptor group galaxies NGC55, NGC300, and NGC7793. The brightest stars are massive red supergiants (RSGs) with $K \\sim 15 - 15.5$. The peak RSG brightness is constant to within $\\sim 0.5$ mag in $K$, suggesting that NGC55, NGC300, and NGC7793 are at comparable distances. Comparisons with bright RSGs in the Magellanic Clouds indicate that the difference in distance modulus with respect to the LMC is $\\Delta \\mu = 7.5$. A rich population of asymptotic giant branch (AGB) stars, which isochrones indicate have ages between 0.1 and 10 Gyr, dominates the $(K, J-K)$ color-magnitude diagram (CMD) of each galaxy. The detection of significant numbers of AGB stars with ages near 10 Gyr indicates that the disks of these galaxies contain an underlying old population. The CMDs and luminosity functions reveal significant galaxy-to-galaxy variations in stellar content. Star-forming activity in the central arcmin of NGC300 has been suppressed for the past Gyr with respect to disk fields at larger radii. Nevertheless, comparisons between fields within each galaxy indicate that star-forming activity during intermediate epochs was coherent on spatial scales of a kpc or more. A large cluster of stars, which isochrones suggest has an age near 100 Myr, is seen in one of the NGC55 fields. The luminosity function of the brightest stars in this cluster is flat, as expected if a linear luminosity-core mass relation is present. ", "introduction": "Recent studies of the structural characteristics of spiral galaxies suggest that the disk and spheroid do not evolve in isolation, but interact throughout the lifetime of a galaxy (e.g. Andredakis, Peletier, \\& Balcells 1995). The observational signatures of these interactions are readily apparent in late-type spiral galaxies. For example, Courteau, de Jong, \\& Broeils (1996) find a scale-free Hubble sequence among late-type spirals, a result which could be explained if the central light concentrations \\footnote{ The presence of a traditional bulge in late-type spirals galaxies has been challenged by Bothun (1992) and Regan \\& Vogt (1994), who concluded that the `bulge' in M33 is actually the central extension of the halo. However, Minniti, Olszewski, \\& Rieke (1993) and Mighell \\& Rich (1995) resolved the innermost regions of M33 into stars, and detected a significant intermediate age population. Minniti {\\it et al.} (1993) argue that, on the basis of stellar content alone, M33 contains a central component that is distinct from the halo. Given this debate, in the current paper the term `central light concentration' is used to refer to what has traditionally been called the `bulge' in late-type spirals.} formed after disks. WFPC1 images discussed by Phillips {\\it et al.} (1996) reveal that the central light concentrations of late-type spirals are structurally distinct from the bulges of early-type spirals, suggesting differences in evolutionary histories. Surveys of the bright stellar content in nearby galaxies provide a direct means of studying the evolution of their central regions. Photometric studies of first ascent and asymptotic giant branch (AGB) populations are of particular interest, as these stars probe evolution during early and intermediate epochs, when the basic properties of the disk and spheroid were being imprinted. While efforts to resolve the central light concentrations of these systems into stars require near-diffraction limited image quality to overcome crowding, the inner disks of many nearby systems can easily be resolved into stars from the ground. Although the Local Group contains the closest, most obvious objects for stellar content surveys, the number of targets is limited to three morphologically diverse spiral galaxies: the Milky-Way, M31, and M33. This limited sample makes it necessary to study more distant systems and, as the nearest collection of galaxies outside the Local Group, the Sculptor Group offers a number of lucrative targets that can be resolved from the ground. In the current paper, deep $J, H$, and $K$ images are used to investigate the photometric properties of cool stars in the inner disks of the Sculptor galaxies NGC55, NGC300, and NGC7793. The morphological types and integrated brightnesses of these galaxies, as assigned by Sandage \\& Tammann (1987), are summarized in Table 1. Throughout this study it is assumed that these galaxies are equidistant with $\\mu_0 = 26.0$, as derived for NGC300 by van den Bergh (1992) from a number of different standard candles. The Galactic reddening towards these galaxies is negligible (Burstein \\& Heiles 1984). There are a number of advantages to conducting photometric surveys of luminous, cool evolved stars at wavelengths longward of $1\\mu$m. Not only is the contrast between bright cool stars and fainter unresolved objects in the disk enhanced at infrared wavelengths, but it is also possible to overcome the effects of line blanketing, which can affect the spectral energy distributions of moderately metal-rich giants at optical wavelengths (e.g. Bica, Barbuy, \\& Ortolani 1991), and complicate efforts to derive bolometric corrections. In addition, near-infrared two-color diagrams can also be used to identify foreground stars, contamination from which may be significant at faint visible magnitudes. NGC55 and NGC300 have been the targets of earlier photometric investigations. Deep broad- and narrow-band surveys of NGC55 (Pritchet {\\it et al.} 1987) and NGC300 (Richer, Pritchet, \\& Crabtree 1985; Zijlstra, Minniti, \\& Brewer 1996) have revealed that the outer disks of these galaxies contain rich AGB populations. A comparison of the AGB luminosity functions suggests that the star-forming histories of NGC55 and NGC300 during intermediate epochs were similar, but not identical (Pritchet {\\it et al.} 1987). Freedman (1984) and Pierre \\& Azzopardi (1988) used $B$ and $V$ photometry to survey the bright young stellar content of NGC300, while Kiszhurno-Kozrey (1988) used CCD observations to construct $(V, B-V)$ CMDs of two fields in NGC55. There is no published photometric study of the stellar content of NGC7793, although Catanzarite {\\it et al.} (1995) report the discovery of Cepheids in this galaxy. The only published infrared survey of these galaxies was carried out by Humphreys \\& Graham (1986), who obtained $JHK$ aperture measurements of red supergiant (RSG) candidates in NGC300. Spectroscopy revealed that almost half of the candidate objects were cool Galactic main sequence stars. The paper is structured as follows. The observations, reduction techniques, and methods used to measure stellar brightnesses are discussed in \\S 2. The luminosity functions (LFs), two-color diagrams (TCDs), and color-magnitude diagrams (CMDs) derived from these data are presented and compared in \\S 3. In \\S 4 the data are used to search for radial population gradients in NGC300 and NGC7793. A summary of the results follows in \\S 5. ", "conclusions": "" }, "9803/astro-ph9803104_arXiv.txt": { "abstract": "We report the results of a 5-GHz southern-hemisphere snapshot VLBI observation of a sample of blazars. The observations were performed with the Southern Hemisphere VLBI Network plus the Shanghai station in 1993 May. Twenty-three flat-spectrum, radio-loud sources were imaged. These are the first VLBI images for 15 of the sources. Eight of the sources are EGRET ($>$~100~MeV) $\\gamma$-ray sources. The milliarcsecond morphology shows a core-jet structure for 12 sources, and a single compact core for the remaining 11. No compact doubles were seen. Compared with other radio images at different epochs and/or different frequencies, 3 core-jet blazars show evidence of bent jets, and there is some evidence for superluminal motion in the cases of 2 blazars. The detailed descriptions for individual blazars are given. This is the second part of a survey: the first part was reported by Shen \\etal (1997). ", "introduction": "Blazar is the collective name for BL Lac objects, optically violent variables and highly polarized quasars, all of which share extreme observational properties that distinguish them from other active galactic nuclei. These properties include strong and rapid variability, high optical polarization, weak emission lines, and compact radio structure (cf. Impey 1992). About 200 blazars have been identified (cf. Burbidge \\& Hewitt 1992). A possible explanation for the blazar phenomenon within a unified scheme for active galactic nuclei is that their emission is beamed by the relativistic motion of the jets traveling in a direction close to the observer's line of sight. This beaming argument is strengthened by the recent CGRO (Compton Gamma Ray Observatory) discovery that most of the detected high-latitude $\\gamma$-ray sources are blazars (e.g. Dondi \\& Ghisellini 1995). A comprehensive theoretical review of these sources has been made by Urry \\& Padovani (1995). Blazars are an important class of active galactic nuclei because they are thought to be sources with relativistic jets seen nearly end-on. Such sources generally have very compact, flat-spectrum radio cores, which are appropriate for VLBI study. Pearson \\& Readhead (1988) have undertaken a survey of a complete sample consisting of 65 strong northern-hemisphere radio sources. They provided the first well-defined morphological classification scheme, based primarily on the large-scale radio structure and radio spectra of the sources. Most surveys to date, however, including the recent Caltech--Jodrell Bank VLBI Surveys (Polatidis \\etal 1995; Thakkar \\etal 1995; Xu \\etal 1995; Taylor \\etal 1994; Henstock \\etal 1995), have been restricted to northern-hemisphere sources. For example, all the confirmed superluminal radio sources, except the well-known equatorial source 3C~279 (1253$-$055), are in the northern sky (Vermeulen \\& Cohen 1994). This reflects the paucity of southern VLBI observations, the notable exceptions being the systematic one-baseline surveys by Preston \\etal (1985) and Morabito \\etal (1986), and the more extensive SHEVE survey (Preston \\etal 1989 and references therein). Since 1992 we have been carrying out a program to address this deficiency, using VLBI at 5 GHz to study southern radio sources. In an earlier paper (Shen \\etal 1997, hereafter Paper I), we reported the results from the first observing session in 1992 November, and presented images of 20 strong sources selected on the basis of their correlated fluxes on intercontinental baselines. In 1993 May we observed a second sample of southern sources, which is the subject of this paper. Section~2 introduces this blazar sample; Section~3 briefly describes the observations and data reduction procedures; Section~4 presents the results; the summary and conclusions are presented in Section~5. Throughout the paper, we define the spectral index, $\\alpha$, by the convention {S$_{\\small \\nu}\\,\\propto\\,{\\nu}^{\\alpha}$}, and assume {H$_0=100$~km~s$^{-1}$~Mpc$^{-1}$} and q$_0=0.5$. ", "conclusions": "In this paper we have defined a sample of southern-hemisphere core-dominated blazars. Of the 24 blazars in the sample, 3 were observed earlier with the same array. The other 21 in the sample and 2 other sources were observed in 1993 May with the Southern VLBI Network plus the Shanghai radio telescope. This is part of the Southern Hemisphere 5-GHz VLBI Survey project, the aim of which is to improve the study of southern extragalactic radio sources (see Paper I). Our study also adds significantly to the number of sources whose structures can be compared on arcsecond (kpc) and milliarcsecond (pc) scales (Table~5). The misalignment of jet-like structures on these scales is an important unsolved problem for the understanding of compact sources. The main conclusions presented in this paper can be summarized as follows: \\begin{enumerate} \\item We have detected and imaged all 23 radio sources, of which 15 are first-epoch VLBI images. These are PKS~0118$-$272, PKS~0332$-$403, PKS~0426$-$380, PKS~0454$-$234, PKS~0823$-$223, PKS~1034$-$293, PKS~1244$-$255, PKS~1514$-$241, PKS~1936$-$155, PKS~1954$-$388, PKS~2005$-$489, PKS~2155$-$152, PKS~2240$-$260, PKS~2243$-$123 and PKS~2355$-$534. \\item Most of the blazars are resolved and display simple morphology, with 12 having core-jet structures and 11 having single-core structures. Observations with increased sensitivity will probably reveal many more core-jet structures (e.g. 2243--123). We have compared our VLBI images with other radio images. Only 3 (PKS~0438$-$436, PKS~0537$-$441 and PKS~1226+023) of the 12 core-jet blazars were found to have curved jets. Superluminal motion was inferred from two-epoch observations for 2 sources (PKS~0208$-$512, PKS~2243$-$123). \\item Eight of these blazars (PKS~0208$-$512, PKS~0454$-$234, PKS~0521$-$365, PKS~0537$-$441, PKS~1127$-$145, PKS~1226+023, PKS~1424$-$418 and PKS~2005$-$489) have been detected at $>100$~MeV $\\gamma$-ray energies. Together with the other 5 EGRET sources observed in 1992 November (Paper I), a total of 13 southern $\\gamma$-ray-loud blazars have now been imaged by our survey project. A systematic study of the VLBI properties of these $\\gamma$-ray blazars and comparison with other non-$\\gamma$-ray sources will improve our understanding of the beaming characteristics in blazars and the properties of EGRET sources. \\end{enumerate}" }, "9803/astro-ph9803332_arXiv.txt": { "abstract": "A quantitative method is presented to compare observed and synthetic colour-magnitude diagrams (CMDs). The method is based on a $\\chi^2$ merit function for a point $(c_i,m_i)$ in the observed CMD, which has a corresponding point in the simulated CMD within $n\\sigma(c_i,m_i)$ of the error ellipse. The $\\chi^2$ merit function is then combined with the Poisson merit function of the points for which no corresponding point was found within the $n\\sigma(c_i,m_i)$ error ellipse boundary.\\hfill\\break Monte-Carlo simulations are presented to demonstrate the diagnostics obtained from the combined ($\\chi^2$, Poisson) merit function through variation of different parameters in the stellar population synthesis tool. The simulations indicate that the merit function can potentially be used to reveal information about the initial mass function. Information about the star formation history of single stellar aggregates, such as open or globular clusters and possibly dwarf galaxies with a dominating stellar population, might not be reliable if one is dealing with a relatively small age range. ", "introduction": "In the last decade the simulation of synthetic Hertzsprung-Russell and colour-magnitude diagrams (hereafter respectively referred to as HRDs and CMDs) has advanced at a rapid pace. It has been applied successfully in various studies of young clusters in the Magellanic Clouds (Chiosi \\etal\\ 1989; Bertelli \\etal\\ 1992; Han \\etal\\ 1996; Mould \\etal\\ 1997; Vallenari \\etal\\ 1990, 1992, and 1994ab), dwarf galaxies in the Local Group (Aparicio{\\muspc\\&\\muspc}Gallart 1995, 1996; Aparicio \\etal\\ 1996, 1997{\\mmuspc}a,b; Ferraro \\etal\\ 1989; Gallart \\etal\\ 1996a{\\to}c; Han \\etal~1997; Tolstoy~1995, 1996; Tosi \\etal~1991), open clusters in our Galaxy (Aparicio \\etal~1990; Carraro \\etal~1993, 1994; Gozzoli \\etal~1996), and the structure of our Galaxy (Bertelli \\etal~1994, 1995, 1996; Ng 1994, 1997{\\mmuspc}a,b; Ng{\\muspc\\&\\muspc}Bertelli 1996{\\mmuspc}a,b; Ng \\etal~1995, 1996{\\mmuspc}a,b, 1997). \\par Generally all studies focus in the first place on matching the morphological structures at different regions in the CMDs (Gallart 1998). In the recent years a good similarity is obtained between the observed and the simulated CMDs. Unfortunately, the best fit is in some cases distinguished by eye. The morphological differences are large enough to do this and the eye is actually guided by a detailed knowledge of stellar evolutionary tracks. However, the stellar population technique has improved considerably and an objective evaluation tool is needed, to distinguish quantitatively one model from another. \\par Bertelli \\etal~(1992,\\muspc1995) and Gallart \\etal~(1996c) defined ratios to distinguish the contribution from different groups of stars. The ratios are defined so that they are sensitive to the age, the strength of the star formation burst and/or the slope of the initial mass function. Vallenari \\etal~(1996{\\mmuspc}a,b) demonstrated that this is a good method to map the spatial progression of the star formation in the Large Magellanic Cloud. Robin \\etal~(1996), Han \\etal~(1997) and Mould \\etal~(1997) use a maximum likelihood method to find the best model parameter(s), while Chen (1996{\\mmuspc}a,b) adopted a multivariate analysis technique. Different models can be quantitatively sampled through Bayesian inference (Tolstoy 1995; Tolstoy{\\muspc\\&\\muspc}Saha 1996) or a chi-squared test (Dolphin 1997). \\par In principle one aims with stellar population synthesis to generate a CMD which is identical to the observed one. The input parameters reveal the evolutionary status of the stellar aggregate under study. To obtain a good similarity between the observed and the simulated CMD one needs to implement in the model in the first place adequately extinction, photometric errors and crowding. It goes without saying that the synthetic population ought to be comparable with the age and metallicity (spread) of the stellar aggregate. Only with a proper choice of these parameters, one can start to study in more detail the stellar initial mass function and the star formation history for an aggregate. \\par This paper describes a quantitative evaluation method based on the combination of a chi-squared and Poisson merit function% \\footnote{The method introduced in Sect.~2 actually mimics closely the procedure used in the `fit by eye' method.}. It allows one to select the best model from a series of models. In the next section this method is explained and Monte-Carlo simulations are made, to display how the diagnostic diagrams of the residual points can be employed. It is demonstrated that the non-fitting, residual points provide a hint about the parameter that needs adjustment in order to improve the model. This paper ends with a discussion about the method and the diagnostic diagrams together with their limitations. \\hfill\\break It is emphasized that this paper deals with a description of a quantitative evaluation method for CMDs. Aspects related with the implementation of an automated CMD fitting program and comparison with simulated or real data sets will not be considered, because they are not relevant for the general validity of the method described in the following section. ", "conclusions": "A method based on the $\\chi^2$ merit function is presented to compare `observed' with synthetic single stellar populations. Monte-Carlo simulations have been performed to display the diagnostic power from a CMD containing the points for which no corresponding synthetic point was found within a reasonable error ellipse. The simulations indicate that the CMDs of residual points might provide hints about model parameters to be improved. The simulations further indicate that one ought to be cautious with the analysis of stellar luminosity functions and that strong hints can be obtained about the shape of the initial mass function." }, "9803/astro-ph9803326_arXiv.txt": { "abstract": "We present WFPC2--HST photometry of the resolved stellar population in the post-starburst galaxy NGC~1569. The color-magnitude diagram (CMD) derived in the F439W and F555W photometric bands contains $\\sim$2800 stars with photometric error $\\leq$ 0.2 mag down to \\mb, \\mv~$\\simeq$~26, and is complete for \\mv~$\\lsim$~23. Adopting the literature distance modulus and reddening, our CMD samples stars more massive than $\\sim$4~\\MSUN, allowing us to study the star formation (SF) history over the last $\\sim$0.15~Gyr. The data are interpreted using theoretical simulations based on stellar evolutionary models. The synthetic diagrams include photometric errors and incompleteness factors. Testing various sets of tracks, we find that the ability of the models to reproduce the observed features in the CMD is strictly related to the shape of the blue loops of the sequences with masses around 5~\\MSUN. The field of NGC~1569 has experienced a global burst of star formation of duration $\\gsim$0.1~Gyr, ending $\\sim$ 5$-$10~Myr ago. During the burst, the SF rate was approximately constant, and, if quiescent periods occurred, they lasted less than $\\sim$10~Myr. The level of the SF rate was very high: for a single-slope initial mass function (IMF) ranging from 0.1 to 120~\\MSUN\\ we find values of 3, 1, and 0.5~\\Myr\\ for $\\alpha = 3$, 2.6, and 2.35 (Salpeter), respectively. When scaled for the surveyed area, these rates are approximately 100 times larger than found in the most active dwarf irregulars in the Local Group. The data are consistent with a Salpeter IMF, though our best models indicate slightly steeper exponents. We discuss the implications of our results in the general context of the evolution of dwarf galaxies. ", "introduction": "Understanding the status and evolution of dwarf galaxies is of crucial importance in current astrophysics and cosmology. They are important ingredients in common scenarios of galaxy formation, either as building blocks or as left-overs of the formation process (Silk 1987). In addition, dwarf galaxies forming at redshift around and below 1 have been suggested to be responsible for the excess of faint blue galaxies seen in deep photometric surveys (Babul $\\&$ Rees 1992; Babul $\\&$ Ferguson 1996). Dwarf galaxies come in three main classes: (i) dwarf spheroidals, which contain population~II stars and have low surface brightness (e.g., Wirth \\& Gallagher 1984); (ii) dwarf irregulars, with various levels of star-formation activity and a high gas mass (Hunter \\& Gallagher 1986); (iii) starbursting dwarfs, including blue compact dwarfs (Thuan 1991), H~II galaxies (Terlevich \\etal\\ 1991), and blue amorphous galaxies (Sandage \\& Brucato 1979). The boundary between dwarf and giant galaxies is usually taken to be around $M_{\\rm{B}} = -16$ mag (Tammann 1994). An intriguing property of dwarf galaxies is the pronounced dichotomy between the two main classes of dwarf spheroidals and dwarf irregulars: most dwarfs are either gas-poor spheroids or gas-rich disk-like systems, with very few transition objects (van den Bergh 1977). The relationship between the two classes is still poorly understood: attempts to unify the complex zoo of dwarf galaxies have not been fully successful (Binggeli 1994). Some effort has been made to understand the interrelationship between individual classes in terms of an evolutionary sequence (e.g., Gallagher, Hunter, \\& Tutukov 1984). Starbursting dwarfs can play a key role: a powerful starburst might strip a gas-rich dwarf irregular of its gas and transform it into a gas-poor dwarf-elliptical (Kormendy 1985; Marlowe 1997). Understanding the star-formation histories in starbursting dwarf galaxies is a prerequisite for understanding evolutionary connections between different object classes (e.g., Gallagher 1996) since during cosmologically brief periods of time their star formation increases dramatically. This has profound implications for the interstellar medium. Stellar winds and supernovae inject energy and may cause a `blow-out', initiating a galactic superwind (Heckman 1995). Numerous examples of galactic superwinds are known from observations (e.g., Heckman, Armus, \\& Miley 1990) and their cosmological implications have been discussed (e.g., De Young \\& Heckman 1994). However, attempts to relate the superwind properties and the star-formation histories in individual galaxies are quite rare and do not go beyond a qualitative level (Heckman \\etal\\ 1990; Leitherer, Robert, \\& Drissen 1992). Yet establishing such a relation is crucial for theoretical models of the superwind hydrodynamics (Suchkov \\etal\\ 1994; Tenorio-Tagle \\& Mu\\~noz-Tu\\~non 1997). NGC~1569 (= UGC~3056 = Arp~210 = VII~Zw~16 = IRAS~4260+6444) has been indicated by Gallagher, Hunter, $\\&$ Tutukov (1984) as an outstanding object in their sample of active star forming galaxies. Indeed, it is a prime candidate for a detailed study of the star-formation history in a starburst galaxy since at $D = 2.2 \\pm 0.6$~Mpc it is the closest starburst galaxy known (Israel 1988a). Adopting the literature reddening $E(B-V)$ = 0.56 (Israel 1988a), photometry of individual stars is feasible with HST down to $M_{V,0}\\simeq -1.5$. This corresponds to stars with mass of about 3~\\Ms\\ in their core-helium burning phase, and consequently to a look-back time of $\\sim$0.4~Gyr (e.g., Schaller \\etal\\ 1992). The properties of NGC~1569 are typical of a dwarf galaxy. With a distance modulus of $(m-M)_0$ = 26.71, its total absolute B magnitude is $M_{B.0} \\sim -17$ (Israel 1988a), which is intermediate between the two Magellanic Clouds. Its total mass and hydrogen content are estimated $M \\simeq 3.3 \\times 10^8$~\\Ms\\ and $M_H \\simeq 1.3 \\times 10^8$~\\Ms, respectively (Israel 1988a). Numerous studies of the chemical composition exist. The published range of oxygen abundances is very narrow: $12 + \\log (O/H) = 8.25$ (Hunter, Gallagher, \\& Rautenkranz 1982), 8.37 (Calzetti, Kinney, \\& Storchi-Bergmann 1994), 8.26 (Devost, Roy, \\& Drissen 1997), 8.29 (Gonz\\'alez-Delgado et al. 1997), and 8.19 (Kobulnicky \\& Skillman 1997; Martin 1997). Searches for chemical composition gradients were done by Devost et al. and Kobulnicky \\& Skillman. No significant evidence for chemical inhomogeneities was found. If [O/Fe]~=~0.0 is assumed, the average metallicity of NGC~1569 is $Z \\simeq 0.25$~\\Zsun, with about 0.2~dex uncertainty. Thus, NGC~1569 is a gas-rich system with SMC-like composition in a relatively early stage of its chemical evolution. The most detailed study of the warm ($\\sim$10$^4$~K) gas of NGC~1569 was done by Waller (1991), who found evidence for a link between the star-formation history and the morphology and kinematics of extended gas structures. The same connection is suggested by hot ($\\sim$10$^7$~K) gas observed with the ROSAT and ASCA satellites (Heckman \\etal\\ 1995; Della Ceca \\etal\\ 1996). The extended X-ray emission is consistent with a starburst driven galactic superwind which could in principle lead to a large-scale disruption of the interstellar medium. Evidence for large numbers of supernovae was also found from the non-thermal radio spectrum (Israel 1988b). One of the most spectacular properties of NGC~1569 are its `super star clusters'. These high-density star clusters were first detected and discussed by Arp \\& Sandage (1985) and Melnick \\& Moles, \\& Terlevich (1985) and later studied by O'Connell, Gallagher, \\& Hunter (1994), Ho \\& Filippenko (1996), De~Marchi \\etal\\ (1997), and Gonz\\'alez-Delgado \\etal\\ (1997). The super star clusters reflect a recent starburst in which at least $10^5$~\\Ms\\ of gas was transformed into stars within about 1~pc. Super star clusters are similar to young Galactic globular clusters observed shortly after their formation (Meurer 1995). Pre-Costar HST studies of the field population were done by O'Connell \\etal\\ (1994) and Vallenari \\& Bomans (1996) (hereinafter VB). According to the latter authors, NGC~1569 experienced a strong starburst until a few Myr ago, when its activity subsided. The repaired HST offers the possibility of significant improvement over these previous studies. We have therefore embarked on an extensive spectroscopic and photometric HST study in an attempt to confront the observational data with the predictions of theoretical models, and derive detailed informations on the star formation (SF) history in NGC~1569. In a first paper (De Marchi \\etal\\ 1997) we discussed the photometry of the SSCs, while in this paper we present the results for the resolved stars in the same field, which are interpreted in terms of the recent SF history via theoretical simulations. The data and their analysis are described in Section~2. The derived color-magnitude diagram and luminosity function are in Section~3. Our theoretical models are described in Section~4. They are compared with the observations in Section~5. In Section~6 we discuss our results and their implications for the evolution of dwarf galaxies, and we present in Section~7 our general conclusions. ", "conclusions": "In this paper we have presented HST WFPC2 F555W and F439W photometry of resolved stars in the dwarf irregular galaxy NGC~1569. About 2800 stars were measured with photometric error smaller than \\sigmada = 0.2\\,mag out of a total of $\\sim$7000 objects detected. The corresponding CMD extends down to \\mb, \\mv~$\\simeq$~26, and is complete for \\mv~$\\lsim$~23. It samples stars more massive than $\\sim$4~\\Ms\\, allowing us to derive the SF history over the last 0.15~Gyr. The interpretation of the data has been performed via theoretical simulations, based on stellar evolutionary tracks. We have considered several sets of stellar tracks from the Geneva and Padova groups, and were able to reach a satisfactory representation of the data adopting either the Geneva tracks with $Z$~=~0.001, or the Padova tracks with $Z$~=~0.004. Since the applicability of these tracks is related to the shape of the blue loops, which is sensitive to details in the input parameters used in the computations, we consider it premature to dismiss alternative sets of evolutionary tracks. The two sets provide a similar picture of the recent SF history in the field of NGC~1569. The galaxy has experienced a general burst of SF, of duration $\\Delta t \\gsim 0.1$~Gyr, at an approximately constant rate. The burst seems to have stopped $\\sim$5~--~10~Myr ago but SF continues in the HII regions and in the SSCs. If this general burst consists of successive episodes, these must have occurred at a similar rate, and be separated by short quiescent periods. Qualitatively, this behavior looks similar to that inferred for dwarf irregulars in the Local Group. Quantitatively, though, the level of the SFR in this recent burst in NGC~1569 is approximately 2 orders of magnitude higher. We find that the Salpeter IMF is consistent with the observed CMD and LF, but our best simulations are characterized by slightly steeper slopes (2.6 -- 3). This seems to disagree with observations of other starbursts which generally indicate a Salpeter slope (Leitherer 1997). The present study, however, refers to the field population of stars less massive than $\\sim$30~\\Ms, whereas most starburst IMFs are derived for the nuclear burst and for dense OB clusters in the mass range above $\\sim$20~\\Ms. Since most of the SF activity in the NGC~1569 field has subsided about 7~--~10~Myr ago, we have no direct information on the IMF of those stars with lifetimes less than this value. The corresponding minimum masses are around 30~\\Ms. The field-star IMF in our Galaxy and in the Magellanic Clouds is steeper (biased against massive stars) than the cluster IMF (Massey, Johnson, $\\&$ DeGioia-Eastwood 1995a; Massey \\etal\\ 1995b), with the cluster value being close to Salpeter. This could be due to a richness effect, where the most massive stars are not observed simply because of their small expected numbers. NGC~1569 may be another case where the most massive stars with masses above $\\sim$30~\\Ms\\ are preferentially formed in clusters, but we emphasize the different mass range sampled by our and that of LMC/SMC studies. Our preferred IMF is quite similar to that derived for field stars of comparable mass in our Galaxy. This is in agreement with the notion of a virtually constant IMF, at least as far as the shape is concerned, and does not support the expectations of flatter slopes in starbursting regions (e.g., Padoan, Nordlund, $\\&$ Jones 1997), or in low metallicity environments, as sometimes invoked to better reproduce the properties of elliptical galaxies (e.g., Vazdekis \\etal\\ 1996). In the past, the SF in NGC~1569 is likely to have proceeded at a substantially lower average rate than in the recent burst. This follows from the estimate of the gas exhaustion timescale, and from the relatively low metallicity of the ISM in this galaxy. This leaves the possibilities of either an approximately constant SFR, or short strong bursts, with long interburst periods. Detailed chemical evolution models are needed to explore quantitatively the possible evolutionary paths for NGC~1569, and this will be the subject of a forthcoming paper. The recent burst of SF is certainly not the first episode in NGC~1569 (see also VB) but may be the last. The velocity field of the ionized gas (Tomita, Ohta, $\\&$ Saito 1994) and energetics arguments support this possibility (Heckman \\etal\\ 1995). If this was the case, in the future NGC~1569 may turn into a dwarf spheroidal galaxy. In addition, its luminosity evolution would be characterized by passive fading, which seems required by the Babul $\\&$ Ferguson (1996) bursting dwarfs model to explain the excess faint galaxy counts. However, it is also possible that the outflowing gas falls back onto the galaxy, triggering a successive burst of SF, an option which has to be explored with the computation of detailed hydrodynamic modelling for bursting dwarf galaxies (D'Ercole $\\&$ Brighenti 1998). In spite of these uncertainties, our results shows that dwarf galaxies are capable of sustaining bursts of SF at large enough rates to be relevant for the interpretation of the faint galaxy counts." }, "9803/astro-ph9803110_arXiv.txt": { "abstract": "The velocity distribution $f(\\bvel)$ of nearby stars is estimated, via a maximum-likelihood algorithm, from the positions and tangential velocities of a kinematically unbiased sample of 14\\,369 stars observed by the \\hip\\ satellite. $f$ shows rich structure in the radial and azimuthal motions, $v_R$ and $v_\\varphi$, but not in the vertical velocity, $v_z$: there are four prominent and many smaller maxima, many of which correspond to well known moving groups. While samples of early-type stars are dominated by these maxima, also up to about a quarter of red main-sequence stars are associated with them. These moving groups are responsible for the vertex deviation measured even for samples of late-type stars; they appear more frequently for ever redder samples; and as a whole they follow an asymmetric-drift relation, in the sense that those only present in red samples predominantly have large $|v_R|$ and lag in $v_\\varphi$ w.r.t.\\ the local standard of rest (LSR). The question arise, how these old moving groups got on their eccentric orbits? A plausible mechanism known from the solar system dynamics which is able to manage a shift in orbit space is sketched. This mechanism involves locking into an orbital resonance; in this respect is intriguing that Oort's constants, as derived from \\hip\\ data, imply a frequency ratio between azimuthal and radial motion of exactly $\\Omega:\\kappa=3:4$. Apart from these moving groups, there is a smooth background distribution, akin to Schwarzschild's ellipsoidal model, with axis ratios $\\sigma_R: \\sigma_\\varphi:\\sigma_z\\approx1:0.6:0.35$. The contours are aligned with the $v_r$ direction, but not w.r.t.\\ the $v_\\varphi$ and $v_z$ axes: the mean $v_z$ increases for stars rotating faster than the LSR. This effect can be explained by the stellar warp of the Galactic disk. If this explanation is correct, the warp's inner edge must not be within the solar circle, while its pattern rotates with frequency $\\gtrsim13\\kmskpc$ retrograde w.r.t.\\ the stellar orbits. ", "introduction": "\\else ", "conclusions": "\\else" }, "9803/astro-ph9803260_arXiv.txt": { "abstract": "There are now about fifty known radio pulsars in binary systems, including at least five in double neutron star binaries. In some cases, the stellar masses can be directly determined from measurements of relativistic orbital effects. In others, only an indirect or statistical estimate of the masses is possible. We review the general problem of mass measurement in radio pulsar binaries, and critically discuss all current estimates of the masses of radio pulsars and their companions. We find that significant constraints exist on the masses of twenty-one radio pulsars, and on five neutron star companions of radio pulsars. All the measurements are consistent with a remarkably narrow underlying gaussian mass distribution, $m=1.35\\pm0.04M_\\odot$. There is no evidence that extensive mass accretion ($\\Delta m\\gsim0.1M_\\odot$) has occurred in these systems. We also show that the observed inclinations of millisecond pulsar binaries are consistent with a random distribution, and thus find no evidence for either alignment or counteralignment of millisecond pulsar magnetic fields. ", "introduction": "Neutron stars have been the subject of considerable theoretical investigation since long before they were discovered as astronomical sources of radio and X-ray emission (\\cite{bz34b,ov39,whe66}). Their properties are determined by the interplay of all four known fundamental forces---electromagnetism, gravitation, and the strong and weak nuclear forces---but neutron stars remain sufficiently simple in their internal structure that realistic stellar modeling can be done. Measurements of their masses and radii (as well as detailed study of their cooling histories and rotational instabilities) provide a unique window on the behavior of matter at densities well above that found in atomic nuclei ($\\rho_{\\rm nuc}\\approx 2.8\\times10^{14}\\mbox{g~cm}^{-3}$). Observations of neutron stars also provide our only current probe of general relativity (GR) in the ``strong-field'' regime, where gravitational self-energy contributes significantly to the stellar mass. The most precisely measured physical parameter of any pulsar is its spin frequency. The frequencies of the fastest observed pulsars (PSR~B1937+21 at 641.9~Hz and B1957+20 at 622.1~Hz) have already been used to set constraints on the nuclear equation of state at high densities (e.g., \\cite{fidp88}) under the assumption that these pulsars are near their maximum (breakup) spin frequency. However, the fastest observed spin frequencies may be limited by complex accretion physics rather than fundamental nuclear and gravitational physics. A quantity more directly useful for comparison with physical theories is the neutron star mass. The basis of most neutron star mass estimates is the analysis of binary motion. Soon after the discovery of the first binary radio pulsar (\\cite{ht75a}), it became clear that the measurement of relativistic orbital effects allowed extremely precise mass estimates. Indeed, the measurement uncertainties in several cases now exceed in precision our knowledge of Newton's constant $G$, requiring masses to be quoted in solar units $GM_\\odot$ rather than kilograms if full accuracy is to be retained. After several recent pulsar surveys, there are now about fifty known binary radio pulsar systems, of which five or six are thought to contain two neutron stars. It is thus possible for the first time to consider compiling a statistically significant sample of neutron star masses. It is our purpose here to provide a general, critical review of all current estimates of stellar masses in radio pulsar binaries. The resulting catalog, with a careful, uniform approach to measurement and systematic uncertainties, should be of value both to those who wish to apply mass measurements to studies of nuclear physics, GR, and stellar evolution, and as a guide to the critical observations for observational pulsar astronomers. We begin with a discussion of known methods for pulsar mass determination (\\S\\ref{sec:methods}), including a new statistical technique for estimating the masses of millisecond pulsars in non-relativistic systems. In \\S\\ref{sec:estimates} we review all known mass estimates, including new data and analysis where possible. Statistical analysis of the available pulsar mass measurements is presented in \\S\\ref{sec:stat}. We summarize in \\S\\ref{sec:summ}. A second paper will consider mass estimates for neutron stars in X-ray binary systems (\\cite[Paper~II]{ct98}). A detailed discussion of the implications of the combined results of this work and Paper~II for studies of supernovae and neutron star formation, mass transfer in binary evolution, the nuclear equation of state, and GR will occur elsewhere (Paper~III). ", "conclusions": "\\label{sec:stat} For a dozen neutron stars, useful mass constraints are available with no assumptions beyond the applicability of the general relativistic equations of orbital motion to binary pulsar systems. Ten of these stars are members of double neutron star binaries. With the possible exception of PSR~B2127+11C, in the globular cluster M15, the pulsar in each system is believed to have undergone a short period of mass accretion during a high-mass X-ray binary phase ($\\Delta m\\sim10^{-3}M_\\odot$, Taam and van den Heuvel 1986). The companion stars have not undergone accretion; their masses most directly preserve information about the initial mass function of neutron stars. Only two ``millisecond'' pulsars, the end products of extended mass transfer in low-mass X-ray binaries, have interesting mass estimates based on GR alone: PSRs~B1802$-$07 and B1855+09. Because such pulsars must accrete $\\sim0.1M_\\odot$ to reach millisecond periods (\\cite{tv86}), and much more ($\\sim0.7M_\\odot$) in some field decay models (e.g., \\cite{vb95a}), obtaining additional mass measurements of millisecond pulsars is of particular interest in testing evolutionary models and in locating the maximum neutron star mass. As noted in \\S\\ref{sec:cmrr}, the $P_b$--$m_2$ relation can be used to estimate the companion mass in recycled binary systems with circular orbits and orbital periods $P_b\\gtrsim 3$~d. There are now thirteen such millisecond ($P_{\\rm spin}<10$~ms) pulsars known, excluding those in globular clusters (where gravitational interactions may have significantly perturbed the orbital parameters since spin-up). In each case, the measured mass function and the inferred companion mass, together with the requirement that $\\sin i<1$, then yields an upper limit on the mass of the pulsar itself. A number of systems in which this upper limit is particularly constraining have been mentioned in \\S\\ref{sec:nswd}. Additional constraints on the neutron star mass in these systems can be derived using statistical arguments, given a prior assumption about the distribution of binary inclinations. The simplest such assumption is that the binaries are randomly oriented on the sky, though biases toward high or low inclinations are possible in some models (\\S\\ref{sec:uniform}). However, as discussed below, we believe there is currently no evidence for such a bias, so for the remainder of this discussion we assume random orbital orientations. For an individual system, we are interested in the probability distribution\\footnote{We adopt the notation that $p(x ; A)$ is the (marginal) probability density for the random variable $x$, where $x$ depends upon the parameter $A$. Also, $p(x|y; A)$ is the conditional probability density for the random variable $x$ for a given value of the random variable $y$ and parameter $A$.} $p(m_1; f, P_b)$ for the neutron star mass $m_1$ given the measured mass function $f$ and binary period $P_b$. We can neglect the measurement uncertainty in $f$ and $P_b$. Then, the probability distribution for $m_1$ can be written schematically as \\begin{equation} p(m_1; f, P_b) = \\int_0^1 d(\\cos i) \\int_{m_{2,{\\rm min}}(P_b)}^{m_{2,{\\rm max}}(P_b)} dm_2\\, p(m_2; P_b)\\, p(\\cos i)\\, p(m_1|m_2,\\cos i; f) , \\end{equation} where $m_1=f^{-1/2}(m_2 \\sin i)^{3/2}-m_2$ is restricted to positive values. We have evaluated this numerically for each system, assuming that $p(\\cos i)$ is uniform between zero and unity and that $p(m_2; P_b)$ is uniformly distributed within the appropriate factor (see \\S\\ref{sec:cmrr}) of the $m_2$ implied by equations (9)--(11). Not surprisingly, the width of the distribution $p(m_1; f, P_b)$ is dominated by the range of allowed $\\cos i$ rather than the uncertainty in $m_2$ for a given $P_b$. For each of the 13 binaries, we have plotted the cumulative distribution CDF$(m_1)=\\int_0^{m_1} p(m_1')\\, dm_1'$ in Figure~\\ref{fig:cumprob}.\\placefigure{fig:cumprob} The median and 68\\% and 95\\% confidence regions for each pulsar mass is given in Table~\\ref{tab:pbm2}. \\placetable{tab:pbm2} Although several of the pulsars have, under the assumptions made, most likely masses well above $2M_\\odot$, some such results are expected even if all the masses are quite low. In fact, in only one case of the 13 pulsars does $1.35M_\\odot$ lie outside the 95\\% central confidence region (J1045$-$4509), and in 6 cases of 13 is $1.35M_\\odot$ excluded at 68\\% confidence, consistent with chance. It is interesting to ask whether a single, simple distribution of neutron star masses is consistent with all of our observational constraints. We considered two models for this question: a Gaussian distribution of masses with mean $\\hat{m}$ and standard deviation $\\sigma$, and a uniform distribution of masses between $m_l$ and $m_u$ ({\\it cf.} Finn 1994). A maximum likelihood analysis was used to estimate the parameters $\\hat{m}$, $\\sigma$, $m_l$, and $m_u$ (assuming a uniform prior distribution for all four parameters). The resulting 68\\% and 95\\% joint confidence limits on $\\hat{m}$ and $\\sigma$ are shown in Figure~\\ref{fig:all26}, and on $m_l$ and $m_h$ in Figure~\\ref{fig:all26uni}. \\placefigure{fig:all26} \\placefigure{fig:all26uni} In each model, the distribution of neutron star masses is remarkably narrow: the maximum likelihood solutions are $\\hat{m}=1.35M_\\odot$ and $\\sigma=0.04M_\\odot$, and $m_l=1.26M_\\odot$ and $m_u=1.45M_\\odot$. Of course, {\\em any} model (even a poor one) will yield maximum likelihood parameters for a given data set. However, it is obvious by inspection that both the Gaussian and uniform distributions for the neutron star mass are good fits to the extremely narrow observed range of neutron star masses in the double neutron star binaries. While it is difficult to quantify the goodness-of-fit for the entire data set, because of the diverse assumptions made in the various mass estimates and the sometimes highly non-gaussian error estimates, we can easily test some neutron star subsamples against the maximum likelihood gaussian model $m_1=1.35\\pm0.04M_\\odot$. For the thirteen neutron-star--white-dwarf binaries, we used a Monte Carlo technique to evaluate the fit quality. For each binary (with its measured $P_b$ and $f$), we simulated a large number of Monte Carlo trials where the neutron star mass $m_1$ was drawn from the maximum likelihood model, $m_2$ was drawn from the appropriate uniform distribution implied by $P_b$, and $\\cos i$ was drawn from a uniform distribution. The Monte Carlo trials were then used to construct the probability distribution for the mass function, and this distribution was used to compute the cumulative probability for the measured mass function, $p(f^\\prime